Biennial Bibliography 2002-2003 College of Forestry

advertisement
2002-2003
Biennial Bibliography
College of Forestry
and the Oregon Forest Research Laboratory
From the Director
Thank you for taking an interest in our work. The Forest Research Laboratory at Oregon State University is
Oregon’s legislatively established and supported center
for forestry-related research. Our mission is to conduct research that provides new knowledge
for enhancing the values of Oregon’s Living Legacy—28 million acres of forested land.
The science-based knowledge we produce is made available by diverse means to provide help
in solving some of the state’s most pressing economic, environmental, and social concerns.
Our research program is closely linked to an outstanding Forestry Extension program that
delivers on-the-ground educational assistance throughout Oregon. Additionally, our Outreach Education program helps promote research results through short courses, workshops,
field tours, and conferences.
One of the most important and wide-ranging ways we share our research is through publication in a variety of written media. The following bibliography lists and briefly describes
papers published by our scientists between July 1, 2002 and December 31, 2003. In order
to provide more timely access, future bibliographies will appear annually on a calendar-year
basis, rather than biennially, as they have in the past.
The publications are arranged in sections according to our five major program areas. The
published findings represent efforts supported by grants from public and private agencies,
state-mandated studies of emerging issues and needs, and unsponsored research reflecting
the interests and expertise of individual scientists.
Many of these publications are available from the Forestry Communications Group (see
following page) or can be viewed at http://fcg.cof.orst.edu/structur/pubs_view.php. Reprints not available from these sources may be requested directly from the author(s).
This annotated bibliography is a testament to the productivity and continuous efforts of
FRL scientists. Forests and forestry are global, so the work done here has implications
around the world. I hope you will be impressed by the quality of work being produced, the
breadth of subjects, and the usefulness of this knowledge in helping protect, sustain, and
optimize the values of our forests.
Research results find application in many areas
as Oregon Forest Research Laboratory scientists
and their cooperators publish their findings. Papers
published between July 1, 2002 and December 31,
2003 are grouped here according to the Oregon
Forest Research Laboratory’s five program areas:
s Forest Regeneration
s Forest Ecology, Culture, and Productivity
s Integrated Protection of Forests and Watersheds
s Evaluation of Forest Uses, Practices, and Policies
s Wood Processing and Product Performance
To Order Publications
Copies of many Oregon Forest Research Laboratory
publications are available from
Forestry Communications Group
Oregon State University
256 Peavy Hall
Corvallis, OR 97331-7401
Phone: (541) 737-4271
FAX: (541) 737-4077
Email: ForestryCommunications@oregonstate.edu
Web site: http://fcg.cof.orst.edu
Please indicate author(s), title, and publication number
if known.
Editing, word processing, design, and layout by Forestry
Communications Group
Table of Contents
Forest Regeneration
4
Forest Ecology, Culture, and Productivity
10
Integrated Protection of Forests and Watersheds
44
Evaluation of Forest Uses, Practices, and Policy
56
Wood Processing and Product Performance
67
Forest Regeneration
and overwinter survival. The second spring, after the
prescribed burn, new germinants dropped dramatically to only three.
Anekonda, TS, MC Lomas, WT Adams, KL Kavanagh, and SN Aitken. 2002. Genetic variation in
drought hardiness of coastal Douglas-fir seedlings
from British Columbia. Canadian Journal of Forest
Research 32: 1701–1716.
Binkley, D, R Senock, and K Cromack, Jr. 2003. Phosphorus limitation on nitrogen fixation by Facaltaria seedlings. Forest Ecology and Management
186: 171–176.
For silviculturists, geneticists, and forest scientists. Thirty-nine full-sib families of coastal Douglas-fir were
studied to determine genetic variation in droughthardiness traits and their correlations with growth
potential and recovery traits. Three moisture regimes
were used: well-watered (control), mild drought,
and moderate drought for the second season, and
well-watered (control), severe drought, and recovery
from moderate drought in the third. Xylem cavitation in the growth ring was 3–4X greater in moderate to severe drought than in the control. Hydraulic
conductivity in xylem of seedlings in severe drought
was 40% lower than in controls. Foliage damage in
severe drought was 33% greater than in moderate
or mild drought. Drought-hardy families potentially
could be identified at the seedling stage. Hardiness
traits showed a strong intercorrelation and may be
controlled by the same genes. Genetic correlations
were not as strong among hardiness traits in different years. Although injury was low, drought reduced
growth the following year. Drought hardiness affected
growth potential very little under the favorable moisture regime.
For silviculturists and tree physiologists. The availability
of phosphorus (P) in the soil often limits the growth
of nitrogen-fixing trees and therefore may also limit
nitrogen (N) fixation rates. Using soils from plantations of either Facaltaria moluccana (which fixes N) or
Eucalyptus saligna (which does not), the authors tested
whether adding N and P fertilizer affected growth
rates and nitrogen fixation. Adding N did not affect
seedling mass, but adding P increased mass by 75%
and more than doubled N fixation. Soil source had
no effect. The authors used a new method for acetylene reduction assays that may allow nondestructive
repeated sampling of nitrogen fixation.
Bishaw, B, DS DeBell, and CA Harrington. 2003.
Patterns of survival, damage, and growth for
western white pine in a 16-year-old spacing trial
in western Washington. Western Journal of Applied
Forestry 18: 35–43.
For silviculturists. White pine blister disease has severely limited the survival of young western white
pine (Pinus monticola) in plantations. The authors
established a spacing trial of this species in the early
1980s in the southern Cascades of Washington. The
seedlings they used had been raised from seeds from
trees certified as rust resistant. Survival averaged
80% 16 years after planting. Although most mortality was associated with the rust, new infections and
mortality diminished markedly between ages 11 and
16, and 71% of the planted trees were free of rust
at age 16. Antler rub damage from elk was substantial, but was overgrown on most trees in 2–4 years.
Berger, C, and DW Gilmore. 2003. Germination
and survival of spruce seedlings following fire in
northwestern Alberta. Northern Journal of Applied
Forestry 20: 45–47.
For forest managers and silviculturists. The authors
sought to determine germination by month and
over-winter survival of spruce seedlings following
a prescribed burn. Seventy percent of all germination occurred in June, and June seedlings had greater
overwinter survival than July or August seedlings.
Depressed microsites had the highest germination
Forest Regeneration
Rub damage appeared to be related to stem diameter,
rather than to spacing. Early growth rates were much
greater than those in older natural stands or in other
trials. Rust-resistant stock of western white pine
might be appropriate for planting in the Douglas-fir
region.
fir allocated more biomass to roots under full light
and limited water and under limited light and water.
Red alder allocated more biomass to the stem under
limited water and light and allocated more biomass to
aboveground components under full light and limited
water. Unlike Douglas-fir, red alder growth responded
negatively to water limitation. Foliar plasticity to
light was greater in the red alder. The results suggest
that red alder and Douglas-fir can coexist under full
light and limiting water situations, and Douglas-fir
can survive better in the understory.
Brandeis, TJ, M Newton, and EC Cole. 2002. Biotic
injuries on conifer seedlings planted in forest understory environments. New Forests 24: 1–14.
For silviculturists. The authors investigated how partial
overstory retention, understory vegetation management, and protective Vexar tubing affected the frequency and severity of biotic injuries in a two-storied
stand underplanted with western redcedar, Douglasfir, grand fir, and western hemlock. The most prevalent source of damage was browsing; deer browsed
over 74% of Douglas-fir and over 36% of western
redcedar seedlings at least once over the 4 years of the
study. Neither the spatial pattern of thinning nor the
density of residual overstory affected browsing frequency. Spraying subplots may have increased browsing frequency slightly, but the resulting reduction
of the adjacent understory vegetation increased the
volume of all seedlings by 13%. Tubing did not substantially affect seedling survival, browsing frequency,
or fourth-year volume. Greater overstory retention
reduced frequency of second flushing. Chafing by deer
and girdling by rodents and other small mammals
began once seedlings surpassed 1 m in height. Essentially all grand fir seedlings were infected by a foliar
fungus.
Cole, E, A Youngblood, M Newton, C Collet, and H
Frochet. 2003. Effect of competing vegetation on
juvenile white spruce (Picea glauca (Moench) Voss)
growth in Alaska. Annals of Forest Science 60:
573–583.
For site managers and silviculturists. The effect of
competing vegetation on survival and juvenile growth
of white spruce was studied on three units in southcentral Alaska and three units in interior Alaska.
Treatments included herbicide site preparation and
release, and minimization of competition for 5 years
(weed-free). By ages 10–11, spruce height and basal
diameter in all units were significantly greater in the
weed-free treatment than in untreated plots. The efficacy of particular treatments on vegetation in a unit
affected results from the other treatments at the different units. For all units, regression analysis showed
that diameter decreased significantly at year 10 or 11
as competitive cover and overtopping increased.
Hibbs, D, B Withrow-Robinson, D Brown, and R
Fletcher. 2003. Hybrid poplar in the Willamette
Valley. Western Journal of Applied Forestry 18:
281–285.
Chan, SS, SR Radosevich, and AT Grotta. 2003.
Effects of contrasting light and soil moisture
availability on the growth and biomass allocation
of Douglas-fir and red alder. Canadian Journal of
Forest Research 33: 106–117.
For pulp and fiber source producers and poplar tree farmers. The authors studied management and growth of
hybrid poplar (Populus spp.) in the Willamette Valley
of western Oregon on poorly drained soils without
irrigation. Growth rates were very good. Simple notill methods were as effective as ripping and tilling
for plantation establishment. Poplar growth was good
with some herbicides but reduced with two commonly
used in site preparation. The good growth appeared to
be associated with deep, extensive root development.
For silviculturists, ecologists, and forest scientists. The
authors examined allocation of growth and biomass
in seedlings of Douglas-fir [Pseudotsuga menziesii
(Mirb.) Franco] and red alder (Alnus rubra Bong.) under different light and water conditions over 3 years.
In every case, alder growth was greater than that of
Douglas-fir. The largest differences in size were at full
sunlight and soil moisture at field capacity. Douglas-
Forest Regeneration
field performance of Douglas-fir transplants. New
Forests 26(3): 263–277.
Howe, GT, SN Aitken, DB Neale, JD Jermstad, NC
Wheeler, TH Chen, and Y Castonguay. 2003.
From genotype to phenotype: Unraveling the complexities of cold adaptation in forest trees. Canadian Journal of Botany 81: 1247–1266.
For nursery personnel and plant physiologists and developmental biologists. Manure, peat, and vermiculite were
incorporated into seedling rearing medium at low and
high rates under two soil moisture regimes. Xylem
water potential, growth of the whole plant, development of root architecture, and field performance of
Douglas-fir seedlings were assessed with and without
fertilization. Fertilization increased height growth
initially but was associated with decreased height and
diameter growth after two seasons. Soil amendments
neither dramatically improved nor negatively affected
the seedling traits measured. Their use may be justified by reduced disease incidence and improved soil
properties.
For geneticists and silviculturists. Temperate and boreal
trees adapt to winter cold through complex genetic,
physiological, and developmental processes. There
are steep genetic clines for cold adaptation traits in
relation to environmental gradients, and population differentiation is generally stronger for cold
adaptation traits. Even though these traits appear to
be under strong natural selection, genetic variation
within populations is high. Genetic control of cold
adaptation traits appears to be complex. The authors
therefore suggest that research should focus on identifying and developing markers for cold adaptation
candidate genes and then uncovering relationships
between phenotype and genotype by multi-locus,
multiallelic techniques. Such methods may ultimately
allow prediction of the performance of genotypes in
breeding programs and enhance understanding of the
evolutionary ecology of forest trees.
Jayawickrama, KJS. 2003. Genetic improvement and
development of western hemlock in Oregon and
Washington: Review and future prospects. Silvae
Genetica 52: 26–35.
For geneticists, forest scientists, and forest managers.
Western hemlock has been genetically improved in
the USA since 1972. Because of the fear of Swiss
needle cast disease in Douglas-fir, support for continued improvement has been growing. In 1992, a cooperative second-generation program was started on the
foundation of several first-generation programs in
Oregon, Washington, and British Columbia. The best
crosses (based on height growth) were planted at nine
sites. An intensively managed 18-year-old hemlock
forest can produce 4.7 million seedlings/ha/year.
Cuttings can be taken from second-generation forests
beginning in 2003. Oregon and Washington foresters currently plant about 9.3 million western hemlock seedlings annually, and about 52% come from
the managed forests. The improved stock can easily
be spread by using rooted cuttings, which will eventually produce their own seed. The second-generation
cutting should gain 22.4% over the base population
at age 15. Western hemlock could probably be selected for other wood properties as well.
Jacobs, DF, RW Rose, and DL Haase. 2003. Development of Douglas-fir seedling root architecture in
response to localized nutrient supply. Canadian
Journal of Forest Research 33: 118–125.
For forest scientists, forest researchers, plant nutritionists,
reforestation professionals and forest managers. Various
rates of controlled-release fertilizer were applied as
a single layer beneath the root system of seedlings of
Douglas-fir planted in pots. Roots were harvested 3
and 6 months later from areas above, directly in, and
below the fertilizer layer. Higher rates of fertilizer
were detrimental to root penetration. Applications
of controlled-release fertilizer (CRF) both increased
growth and decreased root penetration, depending on
fertilizer rate and time since application. This finding demonstrates the need for understanding CRF
technology in order to maximize seedling growth and
minimize adverse effects.
Jacobs, DF, R Rose, DL Haase, and PD Morgan. 2003.
Influence of nursery soil amendments on water
relations, root architectural development, and
Ketchum, JS, and R Rose. 2003. Preventing establishment of exotic shrubs (Cytisus scoparius (L.) Link.
and Cytisus striatus (Hill)) with soil active herbi-
Forest Regeneration
many questions, most focusing on the possible risks
associated with using HTCs. Knowledge still is insufficient to answer most of these questions, and there is
no evidence that HTCs have lowered production costs
or enhanced yields. Finally, many ethical issues still
have not been resolved.
cides (hexazinone, sulfometuron, and metsulfuron). New Forests 25: 83–92.
For silviculturists, vegetation control biologists, logging
site managers, and land managers. A greenhouse study
to evaluate the potential for commonly used forestry
herbicides to control germination success of Scotch
broom and Portuguese broom was conducted. Three
herbicides, hexazinone, sulfometuron, and metsulfuron, were evaluated. Results suggest that preemergence treatments with hexazinone may prevent broom
establishment; sulfometuron and metsulfuron were
not as effective.
McDowell, SCL, and DP Turner. 2002. Reproductive
effort in invasive and non-invasive Rubus. Oecologia 133: 102–111.
For forest scientists and environmental scientists. The
authors studied reproduction in two closely related
species of Rubus, one of which is invasive, by removing flower buds. They quantified the physiological
costs and the total amount of resources allocated to
reproduction. The invasive species (R. discolor) showed
no reproduction-associated change in water stress,
nitrogen concentration ([N]), 13C, or photosynthesis
(A). The noninvasive species (R. ursinus) exhibited
significant water stress and reduced leaf [N] due to
reproduction. These factors reduced diurnal A, wateruse efficiency, and photosynthetic capacity, especially
during fruiting. Estimates of total gross photosynthesis (Agross) showed a greater decline in the noninvasive
species. The invasive species allocated more resources
directly to flowers and fruit but had a significantly
lower reproductive effort, indicating it may minimize
the trade-off of photosynthesis for reproduction.
Lipow, SR, GR Johnson, JB St. Clair, and KJ Jayawickrama. 2003. The role of tree improvement
programs for ex situ gene conservation of coastal
Douglas-fir in the Pacific Northwest. Forest Genetics 10: 113–120.
For forest geneticists and silviculturists. First-generation
tree improvement programs for coastal Douglasfir in the Pacific Northwest include over 4 million
progeny from 33,928 selections planted on 999 test
sites. Second-generation programs have incorporated
nearly 2,000 of those selections. The first-generation
tests provide a repository for potentially important
low-frequency alleles and for variation in quantitative
traits not now under selection. This genetic variation
can be more easily detected in genetic tests and more
rapidly integrated in tree improvement programs than
it can in large in situ populations. In order to retain
the valuable genetic resources within first-generation
genetic tests for a longer time, the authors developed
a method for creating genetic resource outplantings.
Meilan R, K-H Han, C Ma, SP DiFazio, JA Eaton, EA
Hoien, BJ Stanton, RP Crockett, ML Taylor, RR
James, JS Skinner, L Jouanin, G Pilate, and SH
Strauss. 2002. The CP4 transgene provides high
levels of tolerance to Roundup® herbicide in fieldgrown hybrid poplars. Canadian Journal of Forest
Research 32: 967–976.
Martinez-Ghersa, MA, CA Worster, and SR Radosevich. 2003. Concerns a weed scientist might have
about herbicide-tolerant crops: A revisitation.
Weed Technology 17: 202–210.
For geneticists, plant physiologists, and researchers. The
authors tested two genes for imparting tolerance to
glyphosate (the active ingredient in Roundup®), CP4
and GOX, together in hybrid poplars in field studies over 2 years. Ten percent of the transgenic lines
produced showed no foliar damage or reduced growth
after being sprayed with glyphosate at concentrations
above normal commercial rates. GOX was suspected
to cause undesirable side effects, so several lines were
produced with only CP4 inserted. The lines trans-
For biotechnology researchers, ethicists, farmers, reforestation professionals, weed scientists, and general public. The
authors review questions asked about herbicide-tolerant crops (HTCs) over a decade ago. A symposium
was sponsored by the Weed Science Society of America when the technology was just emerging. HTCs have
now become commonplace. The symposium asked
Forest Regeneration
formed with CP4 alone grew significantly better than
those containing both genes and were damaged less by
glyphosate.
on germination of five seral woody competitors in the
Pacific Northwest: Ceanothus integerrimus, C. velutinus,
Rubus parviflorus, R. ursinus, and R. spectabilis. The
herbicides were applied at six rates, and seeds were
checked for germination and development of true
leaves over 9 weeks. Hexazinone reduced germination
and growth of both Ceanothus species and R. parviflorus except at low rates of application and of R. ursinus
at higher rates. Sulfometuron had a small influence
on seed survival but drastically reduced dry weight
of plantlets of all species. Metsulfuron and atrazine
also decreased the dry weight of seedlings in a dosedependent manner. Atrazine reduced survival rate of
both Ceanothus species, and metsulfuron reduced the
survival rate of C. velutinus.
Radosevich, SR, MM Stubbs, and CM Ghersa. 2003.
Plant invasions—Process and patterns. Weed Science 51: 254–259.
For weed scientists and ecologists. The authors looked at
ways to empirically study and accurately predict plant
invasions by analyzing the invasion process. Invasive
plant species are aggressive, with traits such as small
seed size, short juvenile period, ability to escape native predators, persistent seed bank, and young reproductive age. Ecosystem factors are equally important.
Disturbed or spatially open ecosystems seem to be
particularly vulnerable. Soil, climate, and land use
are extrinsic factors responsible for floristic growth
and persistence of invasive species. DNA analysis and
landscape relationships are also important to determine whether a population will spread.
Rose, R, JS Ketchum, C Collet, and H Frochet. 2003.
Interaction of initial seedling diameter, fertilization and weed control on Douglas-fir growth over
the first four years after planting. Annals of Forest
Science 60: 625–635.
Rose, R, and JS Ketchum. 2002. Interaction of vegetation control and fertilization on conifer species
across the Pacific Northwest. Canadian Journal of
Forest Research 32: 136–152.
For silviculturists and nursery personnel. The authors
evaluated the practices of planting larger stock, applying fertilizer, and controlling weeds longer that
are often used to increase growth rate in plantations.
They tested two sizes of planting stock, fertilizer application at planting and in the next year, and controlling weeds for 2 or 3 years. All treatments influenced growth additively, and there were no significant
interactions among treatment levels. Planting larger
initial stock gave the greatest stem volume gains in
the fourth year. Gains from fertilizer application were
short-lived. The extra year of weed control did not
affect volume growth in years 3 or 4 or fourth-year
stem volume.
For forestry professionals. An experiment evaluating three levels of vegetation competition control,
each with two fertilization treatments with complete
slow-release fertilizer, was installed at five sites. Two
of these sites were planted with Douglas-fir in the
Oregon Coast Range, one with ponderosa pine in
eastern Washington, one with western hemlock in
the coastal hemlock zone in Oregon, and one with
coastal redwood in Northern California. Response
to fertilization was less than from weed control and
impacted growth for only the first year, whereas weed
control continued to influence growth the entire 4
years of the study.
Rosner, LS, and JT Harrington. 2003. Optimizing
acid scarification and stratification combinations
for russet buffaloberry seeds. Native Plants Journal
4(2): 81–86.
Rose, R, and JS Ketchum. 2002. The effect of hexazinone, sulfometuron, metsulfuron, and atrazine on
the germination success of selected Ceanothus and
Rubus species. Western Journal of Applied Forestry
17: 194–201.
For vegetation restorationists and those interested in native plant propagation. Acid scarification, cold moist
stratification, or both are generally recommended as
pretreatments in germinating seeds of russet buffaloberry (Shepardia canadensis). The authors tested
combinations of sulfuric acid scarification time (0 or
For forest scientists and silviculturists. The authors
studied the effects of four pre-emergent herbicides
Forest Regeneration
5 minutes) and stratification times (0, 9, 14 weeks).
Stratification was more effective than scarification if
only one treatment was used. The optimal combination of treatments was soaking in acid for 5 minutes
and stratification for 14 weeks. A test to find the
most effective duration of scarification for different
seed lots is proposed.
been shown to affect conifer seedling performance
positively after outplanting, much less is known on
how well ectomycorrhizal fungi establish under nursery conditions followed by outplanting. Rhizopogon
rubescens was inoculated onto the roots of plug-1
ponderosa pine seedling grown in fumigated and
unfumigated bareroot nursery beds. Survival of inoculated seedlings after outplanting on two harsh, dry
sites was significantly higher than that of uninoculated seedlings on both sites. Fumigation did not affect
field survival, and neither inoculation nor fumigation
affected seedling height before or after outplanting.
Foliar phosphorus content was tested at one site and
was higher in inoculated seedlings.
Simard, SW, MD Jones, DM Durall, GD Hope, RJ
Stathers, NS Sorensen, and BJ Zimonick. 2003.
Chemical and mechanical site preparation: Effects
of Pinus contorta growth, physiology, and microsite quality on grassy, steep forest sites in British
Columbia. Canadian Journal of Forest Research 33:
1495–1515.
Vargas-Hernandez, JJ, WT Adams, and DG Joyce.
2003. Quantitative genetic structure of stem form
and branching traits in Douglas-fir seedlings and
implications for early selection. Silvae Genetica 52:
36–44.
For silviculturists and forest scientists. To investigate
interference mechanisms and appropriate site-preparation methods that might lessen the effects of pinegrass (Calamagrostis rubescens Buckl.) on conifers, the
authors prepared two types of sites in British Columbia and monitored them for 9 years. One type was
treated with glyphosate to remove only the pinegrass.
On the other site, both pinegrass and the forest floor
were excavated. Survival of pine (Pinus contorta Dougl.
ex Loud.) in the first 2 years was >97%, compared
with 78% in controls. Growth was greater in the
chemically treated patches than the excavated for the
first 6 years because removing the forest floor also removed or altered elements key to conifer growth. Soil
temperature was increased and frost was reduced the
most on sites that received a larger area of treatment.
For silviculturists and forest scientists. The relationship of stem growth to stem form and branching was
studied in 2-year-old seedlings of open-pollinated
(OP) and full-sib (FS) families of coastal Douglas-fir
in two replicated nursery regimes. The nursery results
were compared with a study of older trees from the
OP family in the field. Heritability of all traits studied except forking was moderate to strong. Field and
nursery trees had similar genetic relationships among
traits. Nursery-field correlations were consistent
among nursery regimes but not usually strong enough
to be useful for early testing purposes. Exceptions
were number of whorls with steep-angled branches
(WSAB), branch length, and branch angle in older
branched trees. Early selection for traits should result
in a predicted 40–50% gain over those selected at an
older age. Early selection for stem growth potential
alone could have unfavorable impacts on WSAB and
stem sinuosity. Branching traits, therefore, should
also be used as selection criteria.
Steinfeld, D, MP Amaranthus, and E Cazares. 2003.
Survival of ponderosa pine (Pinus ponderosa
Dougl. ex Laws.) seedlings outplanted with Rhizopogon mycorrhizae inoculated with spores at the
nursery. Journal of Arboriculture 29: 197–207.
For nursery and forest regeneration specialists and mycologists. Although ectomycorrhizal fungi repeatedly have
Forest Ecology, Culture, and
Productivity
governmental agencies with private organizations to
facilitate research projects by backing them with planning and resources. The NWBC could be used as a
guide for research cooperatives studying other wildlife
species.
Alig, RJ, DM Adams, and BA McCarl. 2002. Projecting impacts of global climate change on the US
forest and agriculture sectors and carbon budgets.
Forest Ecology and Management 169: 3–14.
For ecosystem modelers, forest scientists, global climate
change analysts, and atmospheric scientists. Analyses of
four scenarios of the biological response of forests to
climate change, drawn from a national assessment,
provide information about the economic effects
to the forest and agricultural sectors and the U.S.
carbon budget. Projected changes in climate generally
lead to increased timber inventories in future years,
lower timber prices, and timber producers’ income
that is most at risk. Adjustment mechanisms to
mitigate climate change impacts include interregional
migration of production, substitution in consumption, and altered stand management.
Aubry, KB, JP Hayes, BL Biswell, and BG Marcot.
2003. The ecological role of tree-dwelling mammals in western coniferous forests, pp. 405–443
in Mammal Community Dynamics: Management
and Conservation in the Coniferous Forests of Western North America, CJ Zabel and RG Anthony,
eds. Cambridge University Press, NY.
For wildlife biologists, ecologists, and forestland managers.
Focusing on only a few vertebrate species or attainment of certain vegetative conditions in managing
forests is too simplistic and is unlikely to result in
long-term sustainability. Management decisions
in western coniferous forests might produce better
results if managers consider the ecological roles of
mammals and the webs to which they belong. Comparing the ecological functions that will be affected
by different management structures will allow managers to choose the plan that is most likely to support
ecosystem integrity over the long term.
Arnett, EB. 2003. Advancing science and partnerships for the conservation of bats and their habitats. Wildlife Society Bulletin 31: 2–5.
For conservation biologists, natural resource managers,
and wildlife biologists. Bats make up nearly 25% of
mammalian species, occupy diverse ecological niches,
and often provide major economic benefits. Loss of
bat habitat and declining populations have become
a concern in recent years, and the need for information about their biology and ecology on which to
base management and policy decisions has increased
substantially. This article introduced the papers in a
special section of the Bulletin that focused on advancing science and partnerships for conservation of bats
and their habitats.
Bachelet, D, RP Neilson, T Hickler, RJ Drapek, JM
Lenihan, MT Sykes, B Smith, S Sitch, and K
Thonicke. 2003. Simulating past and future dynamics of natural ecosystems in the United States.
Global Biogeochemical Cycles 17(2): 14-1–14-21.
Arnett, EB, and JB Haufler. 2003. A customer-based
framework for funding priority research on bats
and their habitats. Wildlife Society Bulletin 31:
98–103.
For natural resource managers and wildlife researchers.
The Northwest Bat Cooperative (NWBC) aligns
10
For ecosystem modelers and forest scientists. Two climate change scenarios, one warmer than the other,
were used in DGVMs (Dynamic Global Vegetation
Models) to simulate natural ecosystems as the second phase of the Vegetation/Ecosystem Modeling
and Analysis Project. Both DGVMs included sulfate
aerosols and assumed a gradual CO2 increase. They
simulated reduced southwestern desert areas, westward expansion of eastern deciduous forests, and
Forest Ecology, Culture, and Productivity
expansion of forests in the western part of the Pacific
Northwest and in north-central California. Both predicted an increase in total biomass burnt in the next
century. The two scenarios simulated different carbon
results. Similarities were due to climate forcing; differences were due to structural differences between
the models and sensitivity to CO2. Results are compared with data and a spatial index derived from the
Palmer Drought Severity Index.
For ecologists and wildlife biologists. The diameter at
breast height (dbh) of 700 narrowleaf cottonwood
(Populus angustifolia) and black cottonwood (P. trichocarpa) trees growing in the Lamar Valley (Yellowstone
National Park) was measured. Almost no trees of
either species measured between 5–29 cm dbh; most
were between 30–110 cm. Establishment date for the
narrowleaf cottonwoods were estimated from relationships between age and diameter. These relationships and the dbh data indicated that there has been
very little cottonwood recruitment over approximately
the last 60 years. Fire history, flow regimes, channel
migration, and factors affecting stand development
did not appear to be related to the lack of recruitment, but the beginning of the period coincided with
the elimination of wolves from the park. Loss of this
predator allowed elk to browse riparian plants freely.
Becerra, A, G Daniele, L Dominguez, E Nouhra, and
T Horton. 2002. Ectomycorrhizae between Alnus
acuminate HBK and Naucoria escharoides (Fr.:
Fr.) Kummer from Argentina. Mycorrhiza 12(2):
61–66.
For mycologists and silviculturists. The authors sampled
naturally occurring ectomycorrhizal roots under sporocarps of Naucoria escharoides in natural forest plots at two
sites. They describe the morphology of the ectomycorrhizae, the first detailed description from field specimens.
They also found the internal transcribed spacer region of
the nuclear rDNA, as determined by PCR/RFLP, to be
the same in sporocarps and mycorrhizae.
Bond, B. 2003. Hydrology and ecology meet—and
the meeting is good. Hydrological Processes 17:
2087–2089.
For hydrologists and ecologists. In this commentary
article, the author discusses growth and development of the “hybrid” field, ecohydrology, which studies interactions between the hydrological cycle and
ecosystems. She describes a 10-week seminar series
organized at Oregon State University on “Perspectives on Ecohydrology”. Participants discussed the new
ideas and opportunities offered by the ecohydrology
paradigm, including biodiversity, intra/inter-event
interactions, dimensionality of fluxes, and establishing basin-scale focus. Participants concluded that the
hybrid field has great potential for synergism, so long
as hydrologists and ecologists continue and expand
communication with each other.
Bergen, KM, SG Conard, RA Houghton, ES Kasischke, VI Kharuk, ON Krankina, KJ Ranson,
HH Shugart, AI Sukhinin, and RF Treyfeld.
2003. NASA and Russian scientists observe landcover and land-use change and carbon in Russian
forests. Journal of Forestry 101(4): 34–41.
For forest ecologists and forest resource analysts. Project
teams of the NASA Land-Cover Land-Use Change
Program started cooperating with Russian organizations in 1997 to discover the land-cover and landuse trends of the past, present, and future in Russian
boreal forests. Selected results include information
on forest dynamics, fire and fire behavior, carbon
budgets, and new remote sensing analysis methods.
Our knowledge of the influence of land-cover and
land-use change around the world is growing because
of this and other collaborations with international
organizations and other networks.
Bowling, DR, NG McDowell, BJ Bond, BE Law, and
JR Ehleringer. 2002. 13C content of ecosystem
respiration is linked to precipitation and vapor
pressure deficit. Oecologia 131: 113–124.
For ecosystem scientists. Researchers measured the 13C
content of CO2 respired at night in three ecosystems
along a moisture gradient in Oregon and related the
variation in 13C within each ecosystem to variation
in atmospheric vapor pressure deficit (VPD). The
isotopic composition of respired CO2 was highly
Beschta, RL. 2003. Cottonwoods, elk, and wolves in
the Lamar Valley of Yellowstone National Park.
Ecological Applications 13: 1295–1309.
11
Forest Ecology, Culture, and Productivity
forest tree domestication, pp. 9–44 in Transgenic
Plants: Current Innovations and Future Trends, CN
Stewart, ed. Horizon Scientific Press, Wymondham, England.
correlated with VPD that occurred 3–6 days earlier.
Results indicate that recently fixed carbon cycles
rapidly through ecosystems, and a large proportion of
total ecosystem respiration is derived from recently
fixed carbon.
For tree geneticists, biotechnologists, and physiologists.
Maturation in trees involves numerous programmed
developmental and physiological changes occurring
over decades. These changes are well described, but
little is known about their control mechanisms or how
to alter maturation state. If ability to flower could be
blocked, use of genetically engineered trees in plantations could be ecologically and socially acceptable.
If flowering could be made to occur earlier, breeding
methods now considered impractical in trees would
become feasible. Modifying cambial maturation could
be used to improve wood quality. For many reasons,
Populus promises to be very useful in elucidating the
control mechanisms of maturation and applying that
knowledge successfully.
Brimmer, A, SR Kenny, K Hosaka, JM Trappe, MA
Castellano, JW Spatafora, and W Colgan III.
2002. Monophyly of the Mesophelliaceae. Inoculum 53(3): 22.
For researchers in forest fungus taxonomy and ecology.
This family of truffle-like fungi, important in the
diet of forest mammals in Australia, produces fruiting bodies totally different from any other fungi. Its
relationship to other fungi has been a mystery. Now
DNA analysis shows it to be related to the family
Hysterangiaceae, another truffle-like but world-wide
and common genus.
Brooks, JR, PJ Schulte, BJ Bond, R Coulombe, JC Domec, TM Hinckley, N McDowell, and N Phillips.
2003. Does foliage on the same branch compete
for the same water? Experiments on Douglas-fir
trees. Trees—Structure and Function 17: 101–
108.
Busing, RT, and T Fujimori. 2002. Dynamics of
composition and structure in an old Sequoia sempervirens forest. Journal of Vegetation Science 13:
785–792.
For forest scientists, tree physiologists, and silviculturists.
The gas-exchange responses of sunlit foliage on a
branch were monitored after transpiration of competing foliage on the branch was reduced by bagging and
shading. Responses were compared to these in several
control branches. Several age classes of Douglas-fir
trees were included. Contrary to their expectation,
stomatal conductance (gs) on sunlit foliage did not
increase in any age tree when transpiring leaf area
was reduced, and diurnal changes in water potential,
midday stomatal closure, and photosynthesis did
not differ temporally or in size between treated and
untreated branches. Hydraulic conductance measurements of branch junctions indicate that xylem within
branches is only partially interconnected. Sunlit
foliage on a partially shaded branch apparently has no
advantage with respect to water status over foliage on
a branch in full sun.
For forest scientists and forest ecologists. Over three decades (1972–2001), the dynamics of an old (>1100
years) Sequoia sempervirens forest in northern California were studied with long-term plot data (1.44 ha)
and recent transect data. Changes in the composition
and structure of the tree stratum were minor. Sequoia
mortality and ingrowth were low, and a broad distribution of stem diameters was maintained throughout
the period. Sequoia regeneration was higher at gap
edges. These factors indicate that Sequoia will remain
a dominant species in the study forest.
Busov, VB, R Meilan, DW Pearce, C Ma, SB Rood,
and SH Strauss. 2003. Activation tagging of a
dominant gibberellin catabolism gene (GA 2-oxidase) from poplar that regulates tree stature. Plant
Physiology 132: 1283–1291.
Brunner, AM, B Goldfarb, VB Busov, and SH Strauss.
2003. Controlling maturation and flowering for
12
For tree geneticists, biotechnologists, and physiologists.
A dwarf transgenic hybrid poplar (Populus tremula
× P. albae) was identified. The dwarfing was caused
by overexpression of the major gibberellin catabolic
Forest Ecology, Culture, and Productivity
enzyme and was reversible if gibberellic acid was applied to the shoot apex. Producing semidwarf trees
by transgenic approaches could be quite beneficial,
economically and environmentally.
ences to introduce the available literature on particular species.
Cazares, E, MA Castellano, MP Amaranthus, and
JM Trappe. 2002. Corner’s playground revisited.
Inoculum 53(3): 23.
Campbell, MM, AM Brunner, HM Jones, and S
Strauss. 2003. Forestry’s Fertile Crescent: The
application of biotechnology to forest trees. Plant
Biotechnology Journal 1: 141–154.
For researchers in forest fungus taxonomy and ecology.
EJH Corner, an accomplished British mycologist
confined to Singapore during World War II, spent
much time collecting fungi on the island. The authors
recollected and redescribed many of the truffle-like
fungi he originally discovered there. They comment
on their relationships to other species from Australasia.
For tree geneticists and biotechnologists. Domestication
of crop plants began some 10,000 years ago in the
Fertile Crescent of the Middle East, but it is only in
the last 50 years that forest trees have been domesticated to increase fiber production. Biotechnology
could greatly speed up enhancement of desirable
traits in forest trees. The authors review progress
in applying biotechnology to trees and discuss the
potential for such applications in the future.
Chen, H, and W Hicks. 2003. High asymbiotic N2
fixation rates in woody roots after six years of decomposition: Controls and implications. Basic and
Applied Ecology 4: 479–486.
Castellano, MA, and JM Trappe. 2002. Monophyly,
paraphyly, and generic characterization: Between a
rock and a hard place. Inoculum 53(3): 23.
For ecologists, forest scientists, and those interested in
nutrient recycling. Using acetylene reduction (AR) rates
as a measure, the authors determined asymbiotic N2
fixation associated with decomposing woody roots at
three sites in Oregon. Rates were highest at the HJ
Andrews site in the Cascades, followed by the wettest
site, at Cascade Head, and the driest, at Pringle Falls.
Mean AR rates increased and mass loss decreased
with increasing root size. Root species was not a
factor in AR. Decomposing roots fixed at least four
times more N2 than other asymbiotic sources of N2
fixation. Dead roots can potentially provide a significant amount of nitrogen to forests.
For researchers in forest fungus taxonomy and ecology.
DNA analysis has become an important tool in
understanding relationships among the fungi. It has
produced many surprising results that overturn classical concepts of families and genera. The implications
to fungal nomenclature are discussed with special
reference to potential changes of fungal names.
Castellano, MA, E Cazares, B Fondrick, and T Dreisbach. 2003. Handbook to Additional Fungal Species of Special Concern in the Northwest Forest
Plan. General Technical Report PNW-GTR-572,
USDA Forest Service, Portland OR.
For forest managers and mycologists. The main purpose of this handbook is to help facilitate the survey,
collection, and handling of potential ROD-listed
fungal species by USDA Forest Service and USDI
Bureau of Land Management employees. Each species is represented by a condensed description, a set of
distinguishing features, and information on substrate,
habitat, and seasonality. The authors also present a
list of known sites within the range of the northern
spotted owl, a distribution map, and additional refer-
Cherubini, P, BL Gartner, R Tognetti, OU Braker, W
Schoch, and JL Innes. 2003. Identification, measurement and interpretation of tree rings in woody
species from Mediterranean climates. Biological
Reviews 78: 119–148.
13
For forest scientists and forest ecologists. The authors
reviewed the literature dealing with the effects of
Mediterranean climate, vegetation, phenology, and
ecophysiology on tree ring formation. Studies of tree
rings in Mediterranean regions were scarce. The review found, however, that tree rings are sometimes not
formed because of great spatio-temporal variability of
Forest Ecology, Culture, and Productivity
Mediterranean environmental conditions. In addition, clear seasonality may be lacking, and vegetation
activity may not be always associated with regular
dormancy periods. Difficulties encountered in dating
five species by tree-ring morphology are described,
and the authors propose a classification of tree-ring
formation in Mediterranean climates. Mediterranean
tree rings can be dated and used in dendrochronology,
but sampling sites, species, and sample trees should be
selected with great care.
and effective diffusivity affected the error frequency.
When the applied label was allowed to diffuse for 24
hours, errors decreased by a factor of about three in
the largest aggregate size class.
Cohen, WB, TA Spies, RJ Alig, DR Oetter, TK Maiersperger, and M Fiorella. 2002. Characterizing 23
years (1972–95) of stand replacement disturbance
in western Oregon forests with Landsat imagery.
Ecosystems 5: 122–137.
For landscape ecologists and silviculturists. Patterns of
stand replacement disturbances, such as clearcutting
and wildfire, greatly affect forest ecosystems in western Oregon. The authors contrasted relative amounts
of wildfire and harvest in three major provinces in
the region over 23 years. They also compared harvest
statistics in the principal land ownerships. Nearly
20% of the area was clearcut during the study period,
and wildfire affected 0.7%. Harvest rates of private
industrial landowners, which dominated the Coast
Range province, were about 2.5X greater than those
by public owners. Public ownerships, which dominated the Klamath mounts and Western Cascades
province, had less cutting and more wildfire. Private
industrial owners had larger individual harvest units
that tended to run together over time, whereas public
and private nonindustrial owners tended to cut in
relatively small, spatially dispersed units. Rates of
harvest by all landowners peaked in the late 1980s
and early 1990s and dropped to near the low levels
of the 1970s by the mid-1990s.
Cliff, JB, PJ Bottomley, DJ Gaspar, and DD Myrold.
2002. Exploration of inorganic C and N assimilation by soil microbes with time of flight secondary
ion mass spectrometry. Applied and Environmental
Microbiology 68: 4067–4073.
For researchers in soil nitrogen cycling. A novel application of secondary ion mass spectrometry was used to
detect assimilation of 13C (carbon) and 15N (nitrogen)
by microorganisms at very small (sub-mm) spatial
scales. Fungal hyphae grown on slides contained more
15N in regions influenced by N-rich manure than in
regions influenced by N-deficient straw. TOF-SIMS
is able to locate N-assimilating microorganisms in
soil and quantify the 15N content of cells that have
assimilated 15N-labeled mineral N. It seems promising as a tool with which to explore the factors controlling microsite heterogeneities in soil.
Cliff, JB, PJ Bottomley, R Haggerty, and DD Myrold.
2002. Modeling the effects of diffusion limitations on nitrogen-15 isotope dilution experiments
in soil aggregates. Soil Science Society of America
Journal 66: 1868–1877.
For researchers in soil nitrogen cycling. The authors
examined the effects of mass transfer limitations
on isotope dilution, using spherical diffusion-reaction models. The transport and reaction equations
of NH+4 assumed Fickian diffusion, linear, equilibrium adsorption, zero-order production of natural
abundance 15N, and either pseudo-first-order or
zero-order consumption of NH+4. Pseudo-first-order consumption rate calculations were sensitive to
the adsorption coefficient, but not other transport
parameters. Zero-order consumption and production
rates were consistently underestimated. Aggregate size
Cohen, WB, TK Maiersperger, ST Gower, and DP
Turner. 2003. An improved strategy for regression of biophysical variables and Landsat ETM+
data. Remote Sensing of Environment 84: 561–
571.
14
For researchers interested in applications of remote sensing in ecology. The authors compared three regression
models and their ability to predict LAI for an agroecosystem and live tree canopy cover in a needleleaf
evergreen boreal forest. CCA was used to integrate
multiple dates and SVIs in each of the models. All
the models gave the same coefficients of determination and intercepts but different slopes. The lowest
root mean square error (RMSE) was in the traditional
Forest Ecology, Culture, and Productivity
approach, and the inverse method had the highest
RMSE. Variance of the traditional approach was
lower than that of the observed data set, whereas the
variance of the inverse method was inflated. The reduced major axis (RMA) method had an intermediate
set of predictions for both RMSE and variance.
For researchers in forest soil processes. The allocation of
carbon (C) to underground plant structures is a very
important contributor to C flux in terrestrial ecosystems, but so far has not been well quantified. Previous
work has indicated that the difference between annual
rates of soil respiration and litter fall aboveground
could be used to estimate belowground C allocation.
The authors analyzed new soil respiration and litterfall data, including data from the Ameriflux network.
In general their results agreed with those of earlier
work. Analysis of data from mature forests indicated
that total belowground C allocation is generally much
greater than litterfall. The method, however, has many
limitations. The authors discuss these limitations and
the most appropriate situations for its use.
Compton, JE, MR Church, ST Larned, and WE Hogsett. 2003. Nitrogen export from forested watersheds in the Oregon Coast Range: The role of
N2-fixing red alder. Ecosystems 6: 773–785.
For soil scientists, plant and landscape ecologists, and silviculturists. The authors examined the water chemistry
of 26 small streams in the Coast Range of Oregon in
order to elucidate how red alder influences nutrient
cycling in Pacific Northwest forests. Red alder comprised 94% of basal area of broadleaf cover. Concentrations of nitrate and dissolved organ nitrogen were
positively related to broadleaf cover, more strongly
within entire watersheds than within the riparian
area alone. Annual export of nitrogen varied greatly
among watersheds. Nitrate leaching appeared to
increase cation losses. Their results support the idea
that a single plant species can exert strong control of
ecosystem function, in that leaching from red alder
stands was a major control on nitrogen export from
the watersheds.
Deal, RL, JC Tappeiner, and PE Hennon. 2002.
Developing silvicultural systems based on partial
cutting in western hemlock-Sitka spruce stands of
southeast Alaska. Forestry 75: 425–431.
For forest ecologists, forest managers, and forest resource
researchers. Seventy-three plots in 18 stands that were
harvested 12–96 years ago in southeast Alaska were
evaluated to determine the effects of partial cutting
on species composition, stand structure and growth,
tree size distribution, and tree disease and mortality.
Partially cut stands were diverse and highly complex,
similar to uncut stands. Sitka spruce was maintained
in mixed hemlock-spruce stands over a wide rage of
cutting intensities. There were no significant changes
in tree species composition, stand growth, hemlock
dwarf mistletoe infection and incidence of tree
wounding, or mortality rates with partial cuts. Inclusion of partial cutting in silvicultural systems could
lead to a sustainable timber resource, maintaining
stand structure diversity and old growth characteristics.
Cushing, JB, N Nadkarni, R Dial, and B Bond. 2003.
How trees and forests inform biodiversity and
ecosystem informatics. Computing in Science and
Engineering 5(3): 32–43.
For researchers interested in informatics. The authors
present informatics tools and data infrastructure
techniques developed as part of the Canopy Database
Project. The project combines the skills of computer
scientists and ecologists. The tools developed are useful for helping many researchers integrate their data.
Davidson, EA, K Savage, P Bolstad, DA Clark, PS
Curtis, DS Ellsworth, PJ Hanson, BE Law, Y Luo,
KS Pregitzer, JC Randolph, and D Zak. 2002. Belowground carbon allocation in forest ecosystems
estimated from annual litterfall and IRGA-based
chamber measurements of soil respiration. Agricultural and Forest Meteorology 113: 39–51.
Dean, TJ, SD Roberts, DW Gilmore, DA Maguire,
JN Long, KL O’Hara, and RS Seymour. 2002.
An evaluation of the uniform stress hypothesis
based on stem geometry in selected North American conifers. Trees—Structure and Function 16:
559–568.
15
For silviculturists and forest scientists. In order to test
the uniform stress hypothesis, the authors com-
Forest Ecology, Culture, and Productivity
pared stem taper in 6 conifer species to the taper
expected if stems develop to distribute bending
stress uniformly. They used a model with two fitted
coefficients, δ and ø. Seven of the 12 fitted values
for δ were significantly deviant, 8 were within 10%,
and 11 were within 15% of the hypothesized value
of 0.333. When the values were fitted by relative
height, large portions of the stems were within 10%
of 0.333. Only Pseudotsuga menziesii clearly did
not fit into the uniform-stress hypothesis. For many
of the fitted values, ø was inversely related to the
modulus of elasticity (E) of green wood. Extraordinary values appeared to be accounted for by growing
conditions. E decreased and ø increased with increasing height, suggesting some regulation of bending curvature by adjustments in cross-sectional area.
The results suggest that diameter can be predicted
from bending moment with a simple power function
when E is relatively constant.
acteristics in young, mature, and old-growth ponderosa pine trees. Plant, Cell and Environment 26:
471–483.
For silviculturists, tree physiologists, and wood scientists.
The authors sought to quantify the effects of tree
age and stem position on specific conductivity (ks ),
vulnerability to embolism, and water storage capacity in trunks of young, mature, and old-growth
ponderosa pine. They also sought to determine
relationships between hydraulic characteristics and
radial and height growth rates. Trees of all heights
and ages had 25–60% higher ks in outer sapwood
than inner sapwood. In mature trees only, embolism
started at a lower water potential in the inner sapwood than in the outer. Trees of all ages had similar
capacitances at the top. Increasing ks or vulnerability to embolism increased capacitances sharply. Hydraulic characteristics were correlated with height,
but not diameter, growth rate. High ks was related
to the slowest yearly increase in sapwood area and
a small percentage of latewood. Height growth rate
was associated with high vulnerability to embolism
and high capacitance.
Domec, JC, and BL Gartner. 2002. How do water
transport and water storage differ in coniferous
earlywood and latewood? Journal of Experimental
Botany 53: 2369–2379.
For silviculturists, tree physiologists, and wood scientists.
The authors measured the specific conductivity (ks )
and vulnerability of xylem to embolism on a single
growth ring and in a subset of earlywood and latewood samples within the same ring to determine the
water transport behavior of earlywood and latewood
in the trunk of 21-year-old Douglas-fir. Earlywood
had 90% of the total flow and 11 times the ks of
latewood. Similar to all parts of the wood, earlywood
suffered a 50% loss of conductivity at -2.2 MPa
(P-50). At high trunk water potential, latewood was
more vulnerable to embolism than earlywood. Field
estimates of trunk water potential indicated that
latewood lost 42% of ks , and earlywood lost 16%.
At field trunk water capacity, higher vulnerability to
embolism was associated with greater water storage
capacity. Air seeding through latewood may occur
directly through pores and seal off at lower pressure
than earlywood pores.
Domec, JC, and BL Gartner. 2003. Relationship
between growth rates and xylem hydraulic char-
Dunham, SM, A Kretzer, and ME Pfrender. 2003.
Characterization of Pacific golden chanterelle
(Cantharellus formosus) genet size using co-dominant satellite markers. Molecular Ecology 12:
1607–1618.
For mycologists and molecular geneticists. Genet size
of the Pacific golden chanterelle was determined in
old-growth and in recently thinned and unthinned
young second-growth stands dominated by Douglasfir. Their results indicate that a cryptic chanterelle
species may be included in fruit-body collections
thought to be of C. formosus. Both genetic types had
small genets; variance in genet size was high. The five
loci studied may not have fully resolved genets. Genet
size did not seem to be affected by treatment.
Dunham, SM, TE O’Dell, and R Molina. 2003. Analysis of nrDNA sequences and microsatellite allele
frequencies reveals a cryptic chanterelle species
Cantharellus cascadensis sp. nov. from the American Pacific Northwest. Mycological Research 107:
1163–1177.
16
Forest Ecology, Culture, and Productivity
For mycological systematists and geneticists. In the past,
the yellow chanterelles in the Pacific Northwest have
been called Cantharellus cibarius. The length variability
of the nrDNA internal transcribed spacer, however,
suggests that the C. cibarius-like morphology in fact
covers a species complex. Researchers recently have
identified the chanterelle most frequently harvested
in the Pacific Northwest as C. formosus on the basis of morphological and genetic data. The authors
describe a new species of yellow chanterelle, C. cascadensis sp. nov., characterized by three genetic and
one morphological data set, from the central Cascade
Mountains of Oregon. This species, C. cibarius var.
roseocanus, and the European C. cibarius appear to be
more closely related to white chanterelles (C. subalbidus) than to C. formosus, as indicated by phylogenetic
analyses of the nrDNA large subunit and ITS regions.
Microsatellite data indicated that C. formosus, C.
subalbidis, and C. cascadensis sp. nov. do not interbreed
when they occur together, demonstrating that they are
biological species. The two yellow species cooccuring
in the Pacific Northwest can be separated when fresh
by stipe shape and pileus color.
Erickson, JL, and MJ Adams. 2003. A comparison of
bat activity at low and high elevations in the Black
Hills of western Washington. Northwest Science
77: 126–130.
For zoologists. Differences in activity patterns and
community structure of bats were compared between
low (<150 m) and high (>575 m) elevation sites in
the Capitol State Forest, Washington. Feeding and
total bat activity were higher at the low elevation sites.
Activity of the nonmyotis group was similar at high
and low elevations, whereas activity of the myotis
group was lower at higher elevations. Different activity levels could result from different insect availability, climatic conditions, and morphology of the bat
species.
Erickson, JL, and SD West. 2003. Associations of bats
with local structure and landscape features of forested stands in western Oregon and Washington.
Biological Conservation 109: 95–102.
For ecologists and forest scientists. The authors examined the association of both local structure and landscape context with bat activity in 48 forested stands
in western Oregon and Washington, using ultrasonic
detectors on at least six occasions for each of two
field seasons. Bat activity was negatively associated
with tree density and positively associated with the
standard deviation of tree density and the density of
newly created snags, accounting for 46% of the total
variance in bat activity among stands. Management
of forest-dwelling bats should focus primarily on the
effects of structural attributes at the stand level on
feeding and roosting opportunities.
Emmingham, WH. 2002. Development of ecosystem
management in the Pacific Northwest. Plant Biosystems 136(2): 167–176.
For ecologists and forestland managers. Ecosystem management has influenced forestry practices differently
on federal and private forests. Federal forest managers in western Oregon manage stands for structural
complexity and landscapes to approximate natural
disturbances. Management of private forestlands is
influenced by state and federal regulations and incentives. Within the limits of the regulations, large industrial owners still manage so as to maximize timber
production, with some adopting habitat conservation
plans. Small private owners may balance ecological
requirements, aesthetic preferences, and forest-derived income. Green certification is chosen by some.
The differences in management practices among
ownerships can complement each other. Because the
details of ecosystem management vary by forest type
and management objectives, other regions likely may
require ecosystem management practices different
from those used in the Pacific Northwest.
Falge, E, D Baldocchi, J Tenhunen, M Aubinet, P
Bakwin, P Berbigier, C Bernhofer, G Burba, R
Clement, KJ Davis, JA Elbers, AH Goldstein, A
Grelle, A Granier, J Goumundsson, D Hollinger,
AS Kowalski, G Katul, BE Law, Y Malhi, T Meyers, RK Monson, JW Munger, W Oechel, KT Paw,
K Pilegaard, U Rannik, C Rebmann, A Suyker, R
Valentini, K Wilson, and S Wofsy. 2002. Seasonality of ecosystem respiration and gross primary
production as derived from FLUXNET measurements. Agricultural and Forest Meteorology 113:
53–74.
17
Forest Ecology, Culture, and Productivity
For ecologists, ecosystem modelers, and forest scientists.
Using FLUXNET data, the authors analyzed seasonal patterns of gross primary productivity (F-GPP)
and ecosystem respiration (F-RE) of several different
types of ecosystems. Based on generalized seasonal
patterns, classifications of ecosystems into vegetation
functional types can be evaluated for use in global
productivity and climate change models. Boreal
forests and managed grasslands and crops showed
the greatest seasonal variability. In the temperate
coniferous forests, seasonal patterns of F-GPP and
F-RE were in phase. Temperate deciduous and boreal
coniferous forests had the greatest imbalance between
respiratory and assimilatory fluxes early in the growing season. Annual F-GPP was the lowest in the
boreal climate zone.
Falge, E, J Tenhunen, D Baldocchi, M Aubinet, P
Bakwin, P Berbigier, C Bernhofer, JM Bonnefond, G Burba, R Clement, KJ Davis, JA Elbers,
M Falk, AH Goldstein, A Grelle, A Granier, T
Grundwald, J Gudmundsson, D Hollinger, IA
Janssens, P Keronen, AS Kowalski, G Katul, BE
Law, Y Malhi, T Meyers, RK Monson, E Moors,
JW Munger, W Oechel, KTP U, K Pilegaard, U
Rannik, C Rebmann, A Suyker, H Thorgeirsson,
G Tirone, A Turnipseed, K Wilson, and S Wofsy.
2002. Phase and amplitude of ecosystem carbon
release and uptake potentials as derived from
FLUXNET measurements. Agricultural and Forest
Meteorology 113: 75–95.
For ecosystem modelers and forest scientists. The authors
analyzed seasonal patterns of net ecosystems carbon
exchange (F-NEE) using eddy covariance data of the
FLUXNET data base. Seasonal uptake and release
capacities were in phase in temperate coniferous
forests. Release in temperate deciduous and boreal
coniferous forest was delayed relative to uptake. The
authors created four groupings each for maximum
nighttime release (F-max) and maximum daytime uptake (F-min). The seasonal amplitudes of F-max and
F-min were largest for managed grasslands and crops.
This analysis drew largely on data from Northern
Hemisphere temperate and boreal forest ecosystems.
For regional or global estimates of carbon sequestration potentials, investigations of eddy covariance
should expand in other systems. The results of this
study are potentially important in validating global
carbon cycle modeling.
Garber, SM, and DA Maguire. 2003. Modeling stem
taper of three central Oregon species using nonlinear mixed effects models and autoregressive error
structures. Forest Ecology and Management 179:
507–522.
For tree growth modelers and mensurationists. The
authors developed variable exponent taper models for
Abies grandis, Pinus ponderosa, and Pinus contorta in
the central Oregon Cascades. When two random effects were included, autocorrelation was only partially
reduced. They incorporated a first-order continuous
autoregressive error process into the models in order
to give valid tests of significance on model parameter
estimates. The model was effective for several species
mixes and a range of spacings. Volumes estimated for
P. ponderosa were similar to those from previously developed equations for the east slope of the Cascades,
but estimates differed from east side results for both
A. grandis and P. contorta. Using site-specific volume
and taper equations is important when assessing tree
response to silvicultural treatments.
Garman, SL, JH Cissel, and JH Mayo. 2003. Accelerating Development of Late-successional Conditions in
Young Managed Douglas-fir Stands: A Simulation
Study. General Technical Report PNW-GTR-557,
USDA Forest Service, Pacific Northwest Research
Station, Portland OR.
18
For silviculturists and modelers. Using the
ZELIG>PNW (3.0) model, the authors evaluated
how experimental thinning treatments affected
development of late successional attributes and
extracted merchantable volume in young Douglas-fir
stands in the central Cascades of Oregon. Simulations started with a 40-year-old managed stand and
simulated 64 thinning treatments for four rotation
intervals. An additional 64 experimental thinning
treatments were evaluated under stand conditions of
selected treatment in a subsequent harvest rotation.
In general, results confirmed previous recommendations and showed the potential for a range of thinning
treatments to attain late-successional conditions.
Forest Ecology, Culture, and Productivity
The time required to attain such conditions under
the different treatments would be about the same, but
merchantable volume and long-term stand conditions
would differ among treatments.
Gartner, BL. 2002. Sapwood and inner bark quantities in relation to leaf area and wood density in
Douglas-fir. IAWA Journal 23: 267–285.
For silviculturists and wood scientists. A wide variety of
34-year-old Douglas-fir trees were studied in order
to understand the design criteria for sapwood quantity. Cumulative leaf area had a unique distribution
among factors studied, increasing from the tip to the
base of the crown and then remaining constant. The
ratio of leaf area to sapwood area varied greatly. Sapwood width was fairly constant from tip to stem. This
suggests that sapwood quantity is related to radial gas
diffusion and causes lethal buildup of CO2 or depletion of O2 at the sapwood-heartwood boundary. More
research on actual radial gas diffusion in green wood
is needed.
Gartner, BL, J Roy, and R Huc. 2003. Effects of tension wood on specific conductivity and vulnerability to embolism of Quercus ilex seedlings grown at
two atmospheric CO2 concentrations. Tree Physiology 23: 387–395.
For silviculturists and wood scientists. Seedlings of Quercus ilex L. were grown at normal and elevated CO2 for
16–17 months. Within each group, seedlings were
grown upright or inclined to increase the production
of tension wood (TW) and determine whether this
causes a decrease in hydraulic function. The stem
base, but not the stem tip, formed large amounts of
TW in both CO2 environments. Specific conductivity
(ks) and vulnerability to embolism were not affected
by stem inclination or the amount of TW. Samples
with more TW had higher vessel frequency and wood
density, similar average vessel lumen area and vessel lumen fraction, and more vessels in the smallest
diameter class than samples with less TW. Vessel
frequency, lumen fraction, and wood density were
similar in samples under both CO2 levels, but vessel
lumen area was larger in elevated CO2. The ks values
were lowest at the stem and increased to the tip. Wood
density followed an opposite trend.
Grubisha, L, JM Trappe, R Molina, and J Spatafora.
2002. Biology of the ectomycorrhizal genus
Rhizopogon. VI. Re-examination of infrageneric
relationships inferred from phylogenetic analysis
of ITS sequences. Mycologia 94: 607–619.
For researchers in forest fungus taxonomy and ecology.
DNA analyses of this large genus of mycorrhizal,
truffle-like fungi showed that groupings earlier recognized within the genus are largely correct. Nonetheless, the analyses supported raising several sections
of the genus to subgenus level and transferring many
species from one subgenus to another. This further
refines taxonomy of the genus, which is important
in forest nursery inoculation and as food for native
mammals.
Hallet, JG, MA O’Connell, and CG Maguire. 2003.
Ecological relationships of terrestrial small mammals in western coniferous forests, pp. 120–154
in Mammal Community Dynamics: Management
and Conservation in the Coniferous Forests of Western North America, CJ Zabel and RG Anthony,
eds. Cambridge University Press, NY.
For ecologists, small mammal biologists, and conservationists. Understanding the distribution, abundance, and
ecological influences of small mammals in western
forest ecosystems is relevant to both economic and
conservation concerns. Emphasizing the small mammal species inhabiting the forest floor, the authors
review the taxa and diversity of forest small mammals, population studies, their ecological roles, their
patterns of species richness and abundance, their
habitat associations, and their responses to forest
management. They also suggest directions for future
research.
Hann, DW, and ML Hanus. 2002. Enhanced Diameter-Growth-Rate Equations for Undamaged and
Damaged Trees in Southwest Oregon. Research
Contribution 39, Forest Research Laboratory,
Oregon State University, Corvallis.
19
For silviculturalists and forest managers. Using weighted
nonlinear regression, the authors developed equations
for predicting 5-year diameter-growth rate of eight
conifer and nine hardwood tree species in Oregon.
Forest Ecology, Culture, and Productivity
These equations for damaged and undamaged trees
are being incorporated into ORGANON, a model for
predicting the development of stands for southwest
Oregon. The different effects of specific damaging agents that impact 5-year diameter-growth rate
significantly are explored.
Hann, DW, and ML Hanus. 2002. Enhanced HeightGrowth-Rate Equations for Undamaged and Damaged Trees in Southwest Oregon. Research Contribution 41, Forest Research Laboratory, Oregon
State University, Corvallis.
For silviculturalists and forest managers. Researchers
developed equations for predicting the 5-year heightgrowth rate of six conifer species from southwest
Oregon. These equations are being incorporated into
ORGANON, a model for predicting the development
of stands. They extend the previous model to older
stands and to stands with a heavier component of
hardwood tree species. These equations include a full
characterization of stand development and prediction
of the presence and frequency of agents damaging
trees within the stand and their subsequent impact on
tree attributes such as total height, height to crown
base, and diameter-growth, height-growth, and mortality rates.
tions are presented to aid future modelers, and alternative modeling approaches are explored.
Hansen, K, and JM Trappe. 2002. Terfeziaceae E.
Fisch, in Notes on Ascomycete Systematics, Eriksson, OE, HO Baral, RS Currah, K Hansen, CP
Kurtzman, T Laessøe, and G Rambold, eds. Myconet 8:33–34.
For researchers in forest fungus taxonomy and ecology.
The two genera formerly placed in this truffle family, Mattirolomyces and Terfezia, were shown by DNA
analysis to belong in the family Tuberaceae within the
order Pezizales. Hence, the family name Terfeziaceae
can be discarded.
Hayes, JP, and JC Hagar. 2002. Ecology and management of wildlife and their habitats in the Oregon
Coast Range, pp. 99–134 in Forest and Stream
Management in the Oregon Coast Range, SD
Hobbs, JP Hayes, RL Johnson, GH Reeves, TA
Spies, JC Tappeiner II, and GE Wells, eds. Oregon
State University Press, Corvallis.
For wildlife biologists, ecologists, and forest managers.
More than 200 species of wildlife (defined by the
authors as terrestrial vertebrates) live in the Oregon
Coast Range. This diversity is possible largely because
the Coast Range provides a broad range of natural
habitats. Recently, the influence of land management
on biodiversity has been receiving increased attention on aesthetic, ethical, and utilitarian grounds.
Land management activities affect the availability
of suitable habitat, which is a key influence on the
abundance, distribution, and diversity of wildlife. The
authors discuss habitat components at the stand level,
the special concerns with riparian areas, and the role
of landscape structure with respect to wildlife presence and abundance.
Hann, DW, DD Marshall, and ML Hanus. 2003.
Equations for Predicting Height-To-Crown-Base,
5-Year Diameter-Growth Rate, 5-Year HeightGrowth Rate, 5-Year Mortality Rate, and Maximum Size-Density Trajectory for Douglas-Fir and
Western Hemlock In the Coastal Region of the
Pacific Northwest. Research Contribution 40,
Forest Research Laboratory, Oregon State University, Corvallis.
For forest researchers, forest managers, and silviculturists. Using existing permanent research plot data,
researchers developed equations for predicting
height-to-crown-base, 5-year diameter-growth rate,
5-year height-growth rate, 5-year mortality rate, and
the maximum size-density trajectory for Douglasfir and western hemlock in the coastal region of the
Pacific Northwest. The strengths and weaknesses of
the existing data sets and the modeling and analytical
approaches tested during development of these equa-
Hayes, JP. 2003. Habitat ecology and conservation of
bats in western coniferous forests, pp. 81–119 in
Mammal Community Dynamics: Management and
Conservation in the Coniferous Forests of Western
North America, CJ Zabel and RG Anthony, eds.
Cambridge University Press, NY.
20
For ecologists, wildlife biologists, and conservationists. Six
species of bats that are regularly associated with west-
Forest Ecology, Culture, and Productivity
ern forests and another 13 species that occur in both
forested and nonforested areas of the region may be
affected by forest management. The author reviews
the current status of knowledge about bat ecology,
including roosting ecology, activity areas, riparian
and aquatic habitat, influence of stand area and structure. He notes information gaps and proposes some
directions for future research, and then discusses the
implications of forest management for bat habitat
and conservation.
Hayes, JP, JM Weikel, and MMP Huso. 2003. Response of birds to thinning young Douglas-fir
forests. Ecological Applications 13: 1222–1232.
grazing and elimination of fire, beginning in the
mid-1800s, and displacement of grassland communities by woody vegetation. Their results indicate that
grazing and fire suppression would have reduced soil
organic carbon mass by 17–18% by the 1990s. In
current woodlands, soil and plant carbon stocks were
estimated to be up to 10X higher than they would
have been in the grasslands they replaced. From these
and other results, the authors conclude that woodland
development has increased carbon stocks in the ecosystems significantly in a relative short period.
Hicks, WT, and ME Harmon. 2002. Diffusion and
seasonal dynamics of O2 in woody debris from
the Pacific Northwest, USA. Plant and Soil 243:
67–79.
For forest biologists, forest managers, and ornithologists.
By increasing structural complexity, thinning forest stands may improve habitat for birds and other
vertebrates. The authors investigated the effects of
thinning intensity on diurnal breeding bird populations in western Oregon over 7 years. Detections of
nine species decreased in thinned stands, relative to
controls; eight species increased; and five appeared
unaffected by thinning. In about half of the responsive species, response was greatest in the more heavily
thinned stands. The authors conclude that thinning
dense conifer stands in areas dominated by young
stands increases habitat suitability for some species,
but that some areas should be left unthinned for the
sake of species adversely impacted by thinning.
For forest ecologists. Little is known about oxygen (O2)
levels and diffusion rates in decomposing logs. Using
both laboratory and field studies, the authors studied
how O2 diffusion differed with moisture and density in decayed and sound wood. They also predicted
seasonal changes in O2 concentration in Douglas-fir
and western hemlock logs in an old-growth Douglas
fir forest in the Pacific Northwest and compared them
with observed changes. In laboratory measurements,
the O2 diffusion coefficient increased exponentially
with increased air-filled pore space and decreased
density. The two species in the field did not differ
in O2 concentrations. Seasonal O2 levels were not
consistently related to log moisture, respiration, or
air temperature. Their results indicate that laboratory
measurements of oxygen diffusion may underestimate field oxygen concentrations, perhaps because
of cracks and other passages in decayed logs. Oxygen
probably does not limit decomposition in terrestrial
decayed wood, and anaerobic conditions in logs may
not be as common as has been thought.
Hibbard, KA, DS Schimel, S Archer, DS Ojima, and
W Parton. 2003. Grassland to woodland transitions: Integrating changes in landscape structure
and biogeochemistry. Ecological Applications 13:
911–926.
For forest and grassland scientists, modelers, landscape
ecologists, and biogeochemists. As grazing and fire
regimes have changed under human influence over
the last century, woody vegetation has been replacing grasses in many drylands. The authors linked a
process-based ecosystem model to a transition matrix
model in order to evaluate the size of changes in plant
and soil carbon and nitrogen pools in a subtropical landscape transitioning from grassland to thorn
woodland in Texas. The model included conditions
before Euro-American settlement, followed by heavy
Hicks, WT, ME Harmon, and DD Myrold. 2003. Substrate controls on nitrogen fixation and respiration
in woody debris of the Pacific Northwest, USA.
Forest Ecology and Management 176: 25–35.
21
For ecologists and forest scientists. The authors studied
how nitrogen fixation and respiration in woody debris
are affected by wood species, tissue, and degree of
decay at three sites in the Pacific Northwest. Nitro-
Forest Ecology, Culture, and Productivity
gen fixation and respiration rates were highest in
moderately decayed wood. The warmer, wetter coastal
site had higher actual nitrogen fixation and similar
potential rates than the two interior sites. Bark had
the highest fixation and respiration rates, followed
by sapwood and heartwood, in all the species studied.
Most of the patterns in this study can be explained by
microbial colonization and abundance, resource quality, and climate.
Hobbie, EA, NS Weber, JM Trappe, and GJ van Klinken. 2002. Using radiocarbon to determine the
mycorrhizal status of fungi. New Phytologist 156:
129–136.
For forest researchers, microbiologists, and scientists. 14C
content was measured by accelerator mass spectrometry
in sporocarps, needles, and litter from Woods Creek,
Oregon, along with archived sporocarps. 14C content
of known mycorrhizal fungi resembled that of currentyear needles with an average age of incorporated C of
0–2 years, whereas saprotrophic genera were composed
of C at least 10 years old. 14C could distinguish known
mycorrhizal saprotrophic fungi. 13C measurements
should be interpreted cautiously on species of unknown
status. 14C results for needles and mycorrhizal fungi
suggested that C sources other than atmospheric CO2
may contribute small amounts of C.
Hobbs, SD, and TA Spies. 2002. Introduction, pp.
1–6 in Forest and Stream Management in the
Oregon Coast Range, SD Hobbs, JP Hayes, RL
Johnson, GH Reeves, TA Spies, JC Tappeiner II,
and GE Wells, eds. Oregon State University Press,
Corvallis.
For resource managers, specialists, technicians, and
policymakers. This book summarizes 12 years of
research conducted by scientists affiliated with the
Coastal Oregon Productivity Enhancement (COPE)
Program. It provides new science-based information about the ecology and management of multiple
resources in the Oregon Coast Range.
Horton, TR. 2002. Molecular approaches to ectomycorrhizal diversity studies: Variation in ITS at a
local scale. Plant and Soil 244: 29–39.
For forest scientists, microbiologists, and ecologists.
Sporocarps were collected across a 7-km region of
the Oregon Dunes National Recreation Area. ITSRFLP data are presented for 3–18 sporocarps from
each of 44 ectomycorrhizal species in 18 genera.
Thirty-eight species yielded a single RFLP, and no
two had the same RFLP type. Polymorphic types were
observed in six species. Three species had two types
with one dominating, and three species had three
types with none dominating. These results suggest
that ITS-RFLP data from single samples are robust
for characterizing most of the species at this scale, but
additional analysis should be used.
Hudak, AT, MA Lefsky, WB Cohen, and M Berterretche. 2002. Integration of lidar and Landsat
ETM plus data for estimating and mapping forest
canopy height. Remote Sensing of Environment 82:
397–416.
For resource managers, specialists, technicians, and
policymakers. Key principles related to the ecology and
management of forest and stream resources in the
Oregon Coast Range are described.
Hobbs, SD, JP Hayes, RL Johnson, GH Reeves, TA
Spies, JC Tappeiner II, and GE Wells, eds. 2002.
Forest and Stream Management in the Oregon
Coast Range. Oregon State University Press,
Corvallis.
22
For forest scientists and forest managers. The authors
attempted to integrate lidar and Landsat data to
improve the measurement, mapping, and monitoring
of forest structural attributes. They used five aspatial
and spatial methods for predicting canopy height.
The aspatial regression model results preserved actual
vegetation pattern; however, taller canopies were
underestimated and shorter canopies were overestimated. Spatial models produced less biased results
but reproduced vegetation patterns poorly. Kriged or
cokriged regression residuals were preferable to either
the aspatial or spatial models alone. Sample points at
250-m intervals proved most optimal.
Forest Ecology, Culture, and Productivity
fir trees. Forest Ecology and Management 169:
257–270.
Irvine, J, and BE Law. 2002. Contrasting soil respiration in young and old-growth ponderosa pine
forests. Global Change Biology 8: 1183–1194.
For silviculturists, forest ecologists, and forest scientists.
The authors present a conceptual model of crown
development in Douglas-fir trees, developed by
comparing crown structure among 20-, 40-, and
450-year-old trees. The model examines patterns
of branch volume, branch density, branch diameter,
branch death, and the relationship between mean
branch volume and branch density. Included in the
model are the exponential or general logistic growth
curve, which describes branch volume, and the branch
self-pruning line/curve, which is derived from the
relationship between mean branch volume and branch
density after the onset of branch death. The national
distribution of branch volume shifted toward the upper crown as trees grew older.
For forest scientists and forest researchers. Three years of
specific measurements were taken to determine diel,
seasonal, and interannual patterns of soil efflux in an
old-growth (250 years old) and a recently regenerating (14 years old) ponderosa pine forest in central
Oregon. Interannual variability may be due to soil
moisture availability in the deeper soil horizons on
the old-growth site and the quantity of summer rainfall at the recently regenerating site. Diel mean soil
temperature and soil water potential accounted for
80% of the variance of diel mean soil efflux. Strong
linkages between above- and belowground processes
are suggested by seasonal patterns of soil efflux at
both sites.
Irvine, J, BE Law, and FR Meinzer. 2002. Water
limitations to carbon exchange in old-growth and
young ponderosa pine stands. Tree Physiology 22:
189–196.
For tree physiologists and ecologists. To study how seasonal water deficit influences net ecosystem exchange
of carbon, the authors measured seasonal transpiration; canopy conductance; soil water, temperature,
and respiration; and net ecosystem exchange of
carbon in an old-growth and a recently regenerating
stand of ponderosa pine in Oregon. Both stands had
moderately cold wet winters and hot dry summers.
The seasonal minimum of soil water, reached in August, was the same at both sites. Transpiration rates
were similar in both stands between April and June,
but drought stress then began to increase in trees
in the recently regenerating stand. Predawn water
potential of trees in the young stand also declined
from June through August, whereas that of trees at
the old-growth site remained greater than -0.8 MPa.
Soil water limitations appeared to limit soil respiration and assimilation at the recently regenerating site,
but not at the old-growth site. Net daytime ecosystem
carbon exchange showed no pronounced seasonal pattern at either site between April and November.
Keeley, BW, MB Fenton, and E Arnett. 2003. A
North American partnership for advancing research, education, and management for the conservation of bats and their habitats. Wildlife Society
Bulletin 31: 80–86.
For conservation biologists, natural resource managers,
and wildlife biologists. Conservation of and information about North American bats and their habitats
is needed across the continent. Continent partnerships have successfully promoted the conservation
of other forms of wildlife. The authors discuss the
development, mission and strategic goals, and funding of the North American Bat Conservation Partnership (NABCP). They also recommend priorities
in research, education and management to provide a
framework for coordinating bat conservation continent-wide.
Kellogg, LD, GV Milota, and B Stringham. 2002.
Timber harvesting to enhance multiple resources,
pp. 135–171 in Forest and Stream Management in
the Oregon Coast Range, SD Hobbs, JP Hayes, RL
Johnson, GH Reeves, TA Spies, JC Tappeiner II,
and GE Wells, eds. Oregon State University Press,
Corvallis.
Ishii, H, and N McDowell. 2002. Age-related development of crown structure in coastal Douglas-
For logging planners and forestland managers. Forestland
managers can use timber harvesting to significantly
23
Forest Ecology, Culture, and Productivity
alter the vegetation structure and species composition
of a stand or to place large woody debris in streams.
Such measures can enhance wildlife and fish habitat
or other resource values and help to maintain forest
health. In order to minimize environmental damage,
equipment and harvesting methods should be chosen
carefully for appropriateness to each site. To provide
resource specialists with basic knowledge of forest
operations and harvesting systems, the authors give
an overview of harvesting systems used in the Oregon Coast Range and describe the harvest planning
process. They then discuss timber harvesting in detail
in terms of even-age and uneven-age silviculture and
riparian area management.
that humification of SOM and the stabilization of
soil C was due to microbial changes of organic compounds, rather than to accumulation of recalcitrant
compounds. On the basis of the strong relation they
found between 15N and SOM composition, they
suggest that the degree of SOM humification may be
associated with δ15N values that are higher than those
in fresh litter or other material.
Kratz, TK, LA Deegan, ME Harmon, and WK Lauenroth. 2003. Ecological variability in space and
time: Insights gained from the US LTER program.
BioScience 53: 57–67.
For forest ecologists and researchers. Research from the
Long Term Ecological Research (LTER) Network
provides answers to questions formerly unanswerable and increases our understanding of ecological
phenomena. The authors present examples of this
research ranging widely in spatial scale. Often longterm observations are necessary to discover important
ecological relationships, as shown by the authors’
examples and the LTER experience.
Kennedy, RE, and WB Cohen. 2003. Automated
designation of tie-points for image-to-image coregistration. International Journal of Remote Sensing
24: 3467–3490.
For remote sensing scientists. For image-to-image
registration, common points in both images (image
tie-points, or ITPs) must be identified. The authors
designed and tested software that identifies ITPs by
an automated, area-based technique. The software
was robust to several confounding conditions. Higher
levels of image geometric distortion adversely affected robustness. Over 1600 tests, approximately 8
seconds was required to identify each ITP. In spite of
criticisms of correlation-based techniques such as this
one, the authors feel that they may be useful in many
situations.
Kretzer, AM, DL Luoma, R Molina, and JW Spatafora.
2003. Taxonomy of the Rhizopogon vinicolor species complex based on analysis of ITS sequences
and microsatellite. Mycologia 95: 480–487.
For mycological systematists and geneticists. The authors
used sequence data from the internal-transcribed
spacer region of the nuclear ribosomal repeat and genotypic data from five microsatellite loci to reevaluate species concepts in the Rhizopogon vinicolor species
complex. Both types of analysis separated collections
of this species complex into two distinct clades, suggesting that the five-member complex actually is two
biological species.
Kramer, MG, P Sollins, RS Sletten, and PK Swart.
2003. N isotope fractionation and measures of
organic matter alteration during decomposition.
Ecology 84: 2021–2025.
For forest ecologists and those interested in nutrient
cycling. Because soil organic matter (SOM) is diverse
in its initial chemical composition and stable isotope values, changes that occur in SOM as it decays
and is stabilized are hard to study. The authors used
solid-state 13C Cross Polarization Magic Angle Spin
nuclear magnetic resonance to study stable isotope
ratios with respect to SOM composition in organic
and mineral horizon soil samples from an unpolluted
ecosystem in southeast Alaska. Their results indicate
Kumar, S, KJS Jayawickrama, J Lee, and M Lausberg.
2002. Direct and indirect measures of stiffness
and strength show high heritability in a wind-pollinated radiate pine progeny test in New Zealand.
Silvae Genetica 51: 256–261.
For silviculturists and forest researchers. The authors
sampled 72 first-generation open-pollinated (OP)
families grown in a 12-year-old radiate pine progeny
test to evaluate the effectiveness of several destructive
24
Forest Ecology, Culture, and Productivity
and nondestructive measures of stiffness. By collecting indirect information, stiffness and strength on
clearwood sticks, destructive information on felled
trees, and surrogate (density) information such as
diameter, branching cluster frequency, and straightness, they found that destructive and surrogate information were the best indirect traits for selection of
stiffness and strength. Density appeared to be a better
predictor of strength than of stiffness. Genetic gain
from indirect selection of parents was 78–80% of
that from direct selection. Narrow-sense heritability
estimates and genetic correlations for properties and
traits are provided.
Kurpius, MR, M McKay, and AH Goldstein. 2002.
Annual ozone deposition to a Sierra Nevada ponderosa pine plantation. Atmospheric Environment
36: 4503–4515.
uptake, non-stomatal surface deposition, and gasphase chemistry. Total summer O3 flux was mainly
by gas-phase chemistry; during the winter, stomatal
uptake was dominant. Gas-phase chemistry O3 loss
depended exponentially on temperature. Hydroxyl
radical formation and secondary aerosol growth result
from O3 reacting with biogenically emitted hydrocarbons, with important implications for atmospheric
chemistry and climate.
Kurpius, MR, JA Panek, NT Nikolov, M McKay, and
AH Goldstein. 2003. Partitioning of water flux in
a Sierra Nevada ponderosa pine plantation. Agricultural and Forest Meterology 117 (3–4):173–
192.
For tree physiologists. The cold, wet winters and hot,
dry summers characteristic of the west side of the
Sierra Nevada influence how ponderosa pine trees
in the region partition water between transpiration
and evaporation. The authors measured water fluxes
continually in a young ponderosa pine ecosystem for
1 full year and modeled water fluxes at the site with a
biophysical model, FORFLUX. In summer and fall,
water fluxes were partitioned equally between transpiration and soil evaporation; in winter and spring,
transpiration was dominant. In the dry season, canopy
conductance and transpiration were high in early
morning and mid- to late afternoon and depressed in
midday. The authors suggest that the trees maximize
stomatal conductance and carbon fixation throughout the day during warm, moist periods and in the
morning in dry periods. Model estimates were close to
measured/calculated values during the dry period but
misestimated transpiration and evaporation during
the wet period.
For atmospheric scientists and ecologists. The ozone
concentration and ecosystem scale fluxes above a
ponderosa pine plantation in northern California
were measured continuously from June 1999 to June
2000. While maintaining some activity all year, the
trees were most active during the summer. Cumulative
ozone flux was 127 mmol/m2 over the whole year.
There were high levels of cumulative ozone deposition in non-summer seasons, indicating significant
damage may occur during times when ozone concentrations are not at their highest. Ozone flux had a
closer relation to ozone deposition velocity (O3 Vd )
than to ozone concentration. Controlling variables
had a dynamic relationship to O3 Vd and changed
mostly with water status and phenology. Ozone exposure metrics such as SUM0 were found to be poor
predictors of ozone uptake if periods of ecosystem
stress were not excited.
Laliberte, AS, and WJ Ripple. 2003. Automated wildlife counts from remotely sensed imagery. Wildlife
Society Bulletin 31(2): 362–371.
Kurpius, MR, and AH Goldstein. 2003. Gas-phase
chemistry dominates O3 loss to a forest, implying
a source of aerosols and hydroxyl radicals to the
atmosphere. Geophysical Research Letters 30: art.
no. 1371.
For atmospheric scientists, ecologists, and forest scientists.
The authors partitioned the total measured ecosystem
daytime ozone (O3 ) deposition to a ponderosa pine
forest. The three major loss pathways are stomatal
25
For wildlife and conservation biologists. The authors
assessed the accuracy of counting wildlife from remotely sensed images, using a public domain imageanalysis program. They assessed aerial photos from
hand-held cameras, images from mapping cameras,
and a high-resolution (1 m) satellite image. Count error for the aerial photos ranged from 2.8 to 10.2%.
Forest Ecology, Culture, and Productivity
If spectral and area attributes were combined, single
animals could be separated from groups of 2 or more.
The test with the satellite image was satisfactory, and
61-cm-resolution satellite imagery also should be
suitable. Using image analysis for counting wildlife is
easy and reliable, saves time, and is less labor intensive than manual counts.
For plant ecologists. The authors compared aspen age
structure for elk winter range in Yellowstone National
park with age structure for winter range in two surrounding national forests. Ages were determined
from increment cores taken from three diameter size
classes and three habitat types. The age structure in
Yellowstone differed significantly from that in either
national forest; the national forest age structures were
not different from each other. Since 1930, new overstory stems were recruited in stands in Yellowstone only
in areas of reduced browsing pressure, whereas in the
national forests new stems were recruited in all habitat
types. The authors propose that the patterns observed
are best explained by changes in elk browsing patterns
resulting from difference in predation risk.
Landgren, C, R Fletcher, M Bondi, D Barney, and R
Mahoney. 2003. Growing Christmas Trees in the
Pacific Northwest. Pacific Northwest Extension
Publication 6, Oregon State University, Corvallis.
For current and prospective Christmas tree producers in
Oregon, Washington, and Idaho. The authors present
a primer on growing Christmas trees in the Pacific
Northwest. The potential Christmas tree grower must
take many risks and opportunities into consideration.
A tree farmer can be more successful if he or she
knows the basic tasks required and follows through
with them. A calendar organizes tasks by month. Production characteristics of common Christmas trees in
the region are outlined. Common management and
decision-making options are discussed, including laws
the grower should be aware of.
Landsberg, JJ, RH Waring, and NC Coops. 2003.
Performance of the forest productivity model 3PG applied to a wide range of forest types. Forest
Ecology and Management 172: 199–214.
Law, BE, E Falge, L Gu, DD Baldocchi, P Bakwin, P
Berbigier, K Davis, AJ Dolman, M Falk, JD Fuentes, A Goldstein, A Granier, A Grelle, D Hollinger,
IA Janssens, P Jarvis, NO Jensen, G Katul, Y
Mahli, Matteucci, T Meyers, R Monson, W Munger, W Oechel, R Olson, K Pilegaard, KT Paw, H
Thorgeirsson, R Valentini, S Verman, T Vesala,
K Wilson, and S Wofsy. 2002. Environmental
controls over carbon dioxide and water vapor
exchange of terrestrial vegetation. Agricultural and
Forest Meteorology 113: 97–120.
For forest scientists, forest researchers, and forest managers The authors outlined the 3-PG model and described its calibration procedures. The model can fit
a wide range of forest growth data sets from experimental and commercial plantings with useful accuracy. The authors tested the ability of the model to
predict stand growth at sites representing three levels
of severity that had not been fitted to the data. The
3-PG model can be used with confidence to predict
growth in areas where trees have not been grown and
to explore the effects of environmental conditions on
tree growth and productivity.
For ecologists and forest researchers. This article examined a subset of FLUXNET sites. Net carbon uptake
was larger under diffuse than under direct radiation.
The slope of the relation between monthly gross
ecosystem production and evapotranspiration was
similar among biomes other than tundra. The ratio of
annual ecosystem respiration to gross photosynthesis
was about 0.83, with grasslands having lower values.
Ecosystem respiration and mean annual temperature
were weakly correlated across biomes. Mean annual
temperature and site water balance explained much of
the variation in gross photosynthesis. Leaf area index
over the long-term is limited by water availability. Inter-annual climate changes can limit carbon uptake.
Larsen, EJ, and WJ Ripple. 2003. Aspen age structure
in the northern Yellowstone ecosystem: USA. Forest Ecology and Management 179: 469–482.
Law, BE, A Lindroth, T Meyers, J Moncrieff, R Monson, W Oechel, J Tenhunen, R Valentini, S Verma,
T Vesala, and S Wofsy. 2002. Energy partitioning
26
Forest Ecology, Culture, and Productivity
stored in forests. Lidar-measured canopy structure
and coincident field measurements of above-ground
biomass are compared at sites in the temperate deciduous, temperate coniferous, and boreal coniferous
biomes. Canopy height, canopy cover, and a variety
of canopy-density-weighted heights were estimated.
Mean canopy height followed the expected order, and
nearly all the canopy structure indices were significantly correlated with above-ground biomass. The
exception was canopy cover for temperate deciduous
plots. The authors found that a single equation can
be used to relate canopy structure to above-ground
biomass with no statistically significant bias in predictions for any site.
between latent and sensible heat flux during the
warm season at FLUXNET sites. Water Resources
Research 38: art. no. 1294
For forest researchers, physiologists, and ecologists. The
Bowen ratio [beta (B) = sensible heat flux (H)/latent heat flux (LE)] during the warm season was
investigated at 27 sites over 66 site-years within
FLUXNET. Variability in flux partitioning characteristics was large, but some patterns were detectable.
Deciduous forest sites and the agricultural site had
the lowest surface resistance to water vapor transport
(R-c) and B (0.25–0.50). Coniferous forests had a
larger R-c and B (0.50–1.00 and larger), but R-c
and climatological resistance (R-i) varied from year
to year. Mediterranean climate sites usually had the
highest net radiation, R-c, R-i, and B. Grasslands had
the largest year-to-year variability, depending mostly
on differences in soil water content and R-c.
Law, BE, OJ Sun, J Campbell, S Van Tuyl, and PE
Thornton. 2003. Changes in carbon storage and
fluxes in a chronosequence of ponderosa pine.
Global Change Biology 9: 510–524.
For forest managers, ecologists, physiologists, and silviculturists. The authors estimated carbon storage in vegetation and soil pools, net primary productivity (NPP),
and net ecosystem productivity (NEP) and examined
variation in these ponderosa pine stands of different
ages. NEP was lowest in stands in the initiation age
class (9–23 years), followed by old (190–316 years),
young (56–89 years), and mature (95–106 years)
stands, in order. Similar trends were seen in NPP
without as large a decline in the old stands. Total
ecosystem carbon storage and the ratio of ecosystem
carbon in aboveground wood mass increased up to
150–200 years. Forest inventory data indicate that
a majority of stands are reaching maximum carbon
storage and net carbon uptake.
Lefsky, MA, WB Cohen, DJ Harding, GG Parker, SA
Acker, and ST Gower. 2002. Lidar remote sensing
of above-ground biomass in three biomes. Global
Ecology and Biogeography 11: 393–399.
Lloyd, J, O Shibistova, N Tchebakova, O Kolle, A
Arneth, D Zolotoukhine, JM Styles, and E-D
Schulze. 2002. Seasonal and annual variations in
the photosynthetic productivity and carbon balance of a central Siberian pine forest. Tellus B 54:
590–610.
For researchers in micrometeorology and plant-atmosphere
exchange. Eddy covariance and auxiliary measurements were used to assess seasonal and annual carbon
uptake in a Siberian pine forest. The forest was a
small source of CO2 during the snow season. Photosynthesis commenced as soon as the temperatures
were above freezing, increasing slowly until August
and declining markedly in September. The September
decline may be part of the cold hardening process.
Photosynthetic rate depended mostly on the incoming photon irradiance. In both 1999 and 2000 the
forest was a CO2 sink, at about 13 mol C/m2a. These
data, along with estimates of net primary productivity,
indicate the sink was associated 20% with increasing
plant biomass and 80% with increase in the litter and
soil organic carbon pools.
Lobell, DB, GP Asner, BE Law, and RN Treuhaft.
2002. View angle effects on canopy reflectance
and spectral mixture analysis of coniferous forests
using AVIRIS. International Journal of Remote
Sensing 23: 2247–2262.
For silviculturists, ecologists, and forest scientists. Remote sensing is expected to help in estimating carbon
For landscape ecologists and remote sensing scientists.
Vegetation reflectance depends on sun and sensor
27
Forest Ecology, Culture, and Productivity
geometry. Although this dependence can provide
information about canopy properties, it can also lead
to unmodeled systematic error in single-angle remote
sensing. The authors investigated angular variability of reflectance measurement from the NASA
Airborne Visible/Infrared Imaging Spectrometer
(AVIRIS) and how that variability impacts spectral
mixture analysis (SMA). Reflectance measurements
from AVIRIS can be significantly affected by view
angle, but resulting variability in vegetation cover
estimates appears to be low.
the model predicted maximum LAI satisfactorily.
This may solve the saturation problem in satellite
detection of high forest LAI where the relationship
between NDVI and LAI reaches an asymptote near a
projected LAI value of 5 or 6. Patterns were usually
consistent with satellite NDVI changes. The only exception was monsoon and rain forests in south China,
where satellite detection is extremely difficult. The
authors’ maximum LAI values were generally consistent with global literature data, and they believe that
forest LAI in China is most often >6.
Loescher, HW, SF Oberbauer, HL Gholz, and DB
Clark. 2003. Environmental controls on net
ecosystem-level carbon exchange and productivity
in a Central American tropical wet forest. Global
Change Biology 9: 396–412.
Luoma, DL, JM Trappe, AW Claridge, KM Jacobs, and
E Cazares. 2003. Relationships among fungi and
small mammals in forested ecosystems, pp. 343–
373 in Mammal Community Dynamics: Management and Conservation in the Coniferous Forests of
Western North America, CJ Zabel and RG
Anthony, eds. Cambridge University Press, NY.
For ecologists and atmospheric scientists. Net ecosystem
exchange (NEE) of carbon was estimated from measurements made over 3 years from a 42-m tower in a
tropical wet old-growth forest in Costa Rica with the
eddy-covariance technique. There was no relationship
between NEE and friction velocity (u*) when all the
30-minute averages were used, but a nighttime linear
relationship was found. Mean daytime NEE was -18
mumol CO2/m-2s; mean nighttime NEE was between
4.6 and 7.05 mumol CO2/m-2s, depending on the
method of analysis. Interannual differences in NEE
were apparent, but not seasonal differences. Fifty-one
percent of variation was due to irradiance and 20%
to temperature and vapor pressure deficit together.
Hourly nighttime NEE was weakly related to canopy
air temperature. The forest was a small carbon source
in 1998 and a moderate and strong sink in 1999
and 2000, respectively. The change seems to be due
to the dissipation of warm-phase El Niño effects.
For ecologists, mycologists, and small mammal biologists.
Fungal fruitbodies, especially those of ectomycorrhizal fungi, are an important food source for small
mammals. Drawing on information from western
North America and from Australia, the authors present historical reports of mycophagy, discuss web-oflife relationships and mycophagy, present case studies,
and draw parallels between the northern and southern
hemisphere. The nutritional value of hypogeous sporocarps is discussed. Effects of disturbance on small
mammal mycophagy are treated in some detail, and
the authors suggest directions for future research.
Manning, T, CC Maguire, KM Jacobs, and DL
Luoma. 2003. Additional habitat, diet and range
information for the white-footed vole (Arborimus
albipes). American Midland Naturalist 150(1):
115–122.
Luo, TX, RP Neilson, HQ Tian, CJ Vorosmarty, HZ
Zhu, and SR Liu. 2002. A model for seasonality
and distribution of leaf area index of forests and
its application to China. Journal of Vegetation Science 13: 817–830.
For forest scientists, researchers, and ecologists. The
authors constructed a phenological model of leaf area
index (LAI) of forests, based on biological principles
of leaf growth. Even when maximum LAI was >6,
28
For wildlife biologists and forest managers. Thirteen
white-footed voles were captured in the southern
Cascade Range of Western Oregon, which is a range
extension for the species. Vole encounters were correlated more strongly with basal area of red alder and
percent cover of hazel than with distance to water.
Analysis of fecal samples documented that the voles
were eating hazel pollen from unopened catkins and
also gave the first record of mycophagy from the
Forest Ecology, Culture, and Productivity
species. The association with hazel supports previous
suggestions that the voles are semi-arboreal.
Oregon Coast Range were studied using a plant bioassay. Variations in Frankia populations were found
but did not correlate strongly with any site or soil
characteristics except soil nitrate concentration.
Marshall, H, and G Murphy. 2003. Factors affecting
the accuracy of weighbridge systems. International
Journal of Forest Engineering 14: 67–79.
McDowell, NG, JR Brooks, S Fitzgerald, and BJ Bond.
2003. Carbon isotope discrimination and growth
response of old Pinus ponderosa trees to stand
density reductions. Plant, Cell and Environment
26: 631–644.
For logging operations managers and log scalers. Because
decreasing log size is making individual scaling of
logs both time-consuming and expensive, truck scaling, including the use of weighbridges, is becoming
increasing popular. The authors used interviews with
weighbridge operators, suppliers, and New Zealand
government weights and measures staff and experiments on a small set of weighbridges to identify and,
when possible, quantify the most important factors
affecting weighbridge accuracy. Variation arises from
mechanical, environmental, truck, human, and system-related factors. The combined sources of variation could be as much as 4% of payload weight.
For researchers in forest science and forest managers. To
find out whether old-growth ponderosa pine trees can
respond to thinning, the authors compared ring width
and carbon isotopes in annual rings of old-growth
and control trees. Carbon isotope discrimination (Δ)
increased by 0.89% by the sixth year after thinning.
The unthinned trees had no change in Δ. Basal Area
Increment (BAI) doubled or tripled after thinning,
peaking 1 year after Δ peaked. Stomatal conductance
(g) increased by 25% and photosynthesis increased
by 15%. Pre-dawn leaf water potential was on average 0.11 MPa less negative in the thinned stands.
The results suggest that water availability and g affect
significantly the carbon assimilation and growth of
old trees.
Martin, KJ, and BC McComb. 2003. Amphibian habitat associations at patch and landscape scales in
the central Oregon Coast Range. Journal of Wildlife
Management 64: 672–683.
For ecologists and wildlife biologists. The authors wanted
to clarify multiscale habitat associations with respect to
amphibians. They compared capture rates of 5 species
(26–79 captures) in 7 vegetation patch types and 4
species in 11 patch types to expected capture rates. In 8
species, capture rates were higher in conifer and mixed
large sawtimber than in other patch types. Percent area
of these patch types was associated with capture rates
of 5 species and pattern metrics, with capture rates of
4 species. Because most species were associated with
unique habitats, managing across species would be
difficult. Amphibian abundance and distribution seem
to be strongly associated with mature forest area, patch
richness, pattern, and composition.
McDowell, NG, N Phillips, C Lunch, BJ Bond, and
MG Ryan. 2002. An investigation of hydraulic
limitation and compensation in large, old Douglasfir trees. Tree Physiology 22: 763–774.
Martin, KJ, NJ Ritchie, and DD Myrold. 2003. Nodulation potential of soils from red alder stands
covering a wide age range. Plant and Soil 254:
187–192.
For researchers in symbiotic nitrogen fixation. Populations of Frankia in soils from red alder stands in the
29
For tree physiologists. The authors compared Douglasfir trees in stands grouped by height and approximate
age to test the hydraulic limitation hypothesis. As
tree height decreased, so did carbon isotope discrimination (Δ). Hydraulic conductance decreased
as height increased, especially from 15 m to 32 m.
Photosynthesis (A) and stomatal conductance (gs )
did not decrease with height during summer drought,
but hydraulic conductance (kl) and growth efficiency
did decrease. There may be temporal changes in the
response of gas exchange to height-related changes in
kl, as indicated by the differences between the trend
in gs and A and that in kl and D. They also could be
the result of measurement inadequacies. Sensitivity
analyses showed that kl would have been reduced by
Forest Ecology, Culture, and Productivity
America: A critical review of habitat studies. Wildlife Society Bulletin 31: 30–44.
over 70% in 60-m trees compared with the 15-m
trees, but the actual decrease was 44%.
For conservation biologists, natural resource managers,
and wildlife biologists. Land managers often are asked
to consider habitat needs of bats in their planning.
The authors reviewed peer-reviewed manuscripts appearing from 1960 through 2001. They found that
data about bat habitat were limited in many aspects.
They make several recommendations for improving
the quality of studies on bat habitat and suggest the
development of cooperative research efforts.
McGraw, R, N Duncan, and E Cazares. 2002. Dietary
components of the blue-gray taildropper slug (Prophysaon coeruleum) and the papillose taildropper
slug (Prophysaon dubium) based on fecal analysis.
The Veliger 45: 261–264.
For forestland managers, ecologists, and malacologists.
The blue-gray taildropper (Prophysaon coeruleum) and
the papillose taildropper (P. dubium) are listed as Survey and Manage species in the Northwest Forest Plan,
in part because little is known about their natural history and ecology. Field observations of the two species
suggested that they are mycophagous. Examination of
their fecal pellets showed that both species consume
mushrooms and truffles, roots and other vascular
plant tissue, lichens, and imperfect fungi. Fungi were
the most common items found in samples from both
species. Most of the fungal spores were from hypogeous mycorrhizal taxa. Root tissue may have been
consumed incidentally to feeding on fungal hyphae.
Slugs may be important vectors of fungal spores and
hyphae in the forest.
Mintie, AT, RS Heichen, K Cromack, Jr, DD Myrold,
and PJ Bottomley. 2003. Ammonia-oxidizing
bacteria along meadow-to-forest transects in the
Oregon Cascade mountains. Applied and Environmental Microbiology 69: 3129–3136.
For researchers in soil microbial ecology. This study
combines classic measures of nitrification potential
and modern molecular approaches to studying the
composition of the ammonia oxidizer community in
soils. Although nitrification activities in soils from
adjacent meadows and forests at the HJ Andrews
Experimental Forest were markedly different, the
composition of the ammonia-oxidizing communities
was fairly similar. All ammonia oxidizers were members of the Nitrosospira. Novel strains were isolated
in pure culture.
Meinzer, FC, SA James, G Goldstein, and D Woodruff.
2003. Whole-tree water transport scales with
sapwood capacities in tropical forest canopy trees.
Plant, Cell and Environment 26: 1147–1155.
For tree physiologists and tropical forest scientists. Sapwood capacitance was calculated for four tropical forest canopy trees representing a range of wood density,
tree size, architecture, and taxonomy and used to scale
whole-tree water transport properties. Capacitance
was strongly correlated with minimum branch water
potential, soil-to-branch hydraulic conductance, daily
utilization of stored water, and axial and radial movement of deuterated water used as a tracer. Several of
the transport properties scaled with sapwood capacitance in a species-independent manner, indicating
that this simple variable related to a property of water
transport tissue could reveal convergence in plant
function at several levels of biological organization.
Miller, DA, EB Arnett, and MJ Lacki. 2003. Habitat
management for forest-roosting bats of North
Myrold, DD, and K Huss-Danell. 2003. Alder and
lupine enhance nitrogen cycling in a degraded
forest soil in northern Sweden. Plant and Soil
254: 47–56.
For researchers in symbiotic nitrogen fixation. Natural
variations in δ15N were used to evaluate the inputs of
N from N2-fixation from alders and lupines planted
on a depauperate site in northern Sweden. The presence of N2-fixing plants increases measures of soil
nitrogen availability.
Neilson, RP. 2003. The importance of precipitation
seasonality in controlling vegetation distribution,
pp. 47–71 in Changing Precipitation Regimes
and Terrestrial Ecosystems, JF Weltzin and GR
McPherson, eds. University of Arizona Press,
Tucson.
30
Forest Ecology, Culture, and Productivity
For plant ecologists and biogeographers. After summarizing the broad patterns of climate and vegetation
distribution in North America, the author develops
and examines hypotheses about the importance of
precipitation seasonality in controlling the distribution of vegetation, using several regions of the conterminous United States as case studies. He examines
scenarios of climate of vegetation change and discusses current uncertainties and future directions.
O’Connell, KEB, ST Gower, and JM Norman. 2003.
Net ecosystem production of two contrasting boreal black spruce forest communities. Ecosystems
6: 248–260.
For ecologists and forest scientists. The carbon (C)
budget of a closed-canopy black spruce forest with
feathermoss ground cover (BSFM) on moderately
drained soils was contrasted with that of a similarly
aged open-canopy black spruce forest with sphagnum
ground cover (BSSP) on poorly drained soils. Total C
content did not differ significantly, and both forests
were C sources to the atmosphere. Annual heterotrophic respiration and flux were greater and net ecosystem production was more negative in the BSFM
forest. A lower decomposition rate was indicated in
the BSSP forest. Small differences in edaphic conditions apparently influence ecosystem C accumulation.
Nouhra, E, MA Castellano, and JM Trappe. 2002.
NATS Truffle and truffle-like fungi 9: Gastroboletus molinai sp. nov. (Boletaceae, Basidiomycota),
with a revised key to the species of Gastroboletus.
Mycotaxon 83: 409–414.
For researchers in forest fungus taxonomy and ecology.
The fungal genus Gastroboletus is more abundant in
Oregon and California than anywhere else in the
world. This new species, named in honor of its collector, Dr. Randy Molina, is another restricted to those
states. As mycorrhizal fungi derived from the mushroom genus Boletus, they are particularly interesting
in what they reveal about evolution of fleshy fungi.
Parmenter, AW, A Hansen, RE Kennedy, W Cohen,
U Langner, and R Lawrence. 2003. Land use and
land cover change in the Greater Yellowstone
Ecosystem: 1975–1995. Ecological Applications
13: 687–703.
O’Connell, KEB, ST Gower, and JM Norman. 2003.
Comparison of net primary production and lightuse dynamics of two boreal black spruce forest
communities. Ecosystems 6: 236–247.
For ecologists and forest scientists. The authors compared
net primary production (NPP) and light-use efficiency
(LUE) of two black spruce communities: open-canopy
overstory with sphagnum ground cover (BSSP) and
closed-canopy overstory with feathermoss ground cover
(BSSP). They also quantified the contribution of vegetation layers to total NPP and LUE in a mature black
spruce stand in Saskatchewan, Canada. Total NPP and
ecosystem LUE were significantly greater in the BSFM
community. The communities differed in the contribution of the vegetation layers to NPP and LUE. The
authors conclude that (1) accurate estimation of NPP
and LUE in boreal black spruce ecosystems requires
measurement of all vegetation components and (2) the
accuracy of LUE models is increased by community
specific LUE coefficients.
31
For forest ecosystem modelers, ecologists, wildlife biologists, and forest scientists. The authors quantified the
trajectories and rates of change in land cover and use
across the Greater Yellowstone Ecosystem from 1975
to 1995. Spectral and geographic variables were used
in classification tree regression analysis (CART) to
form functions that created maps using data from
1995, 1985, and 1975. The greatest changes
among the maps were increases in burned, urban,
and mixed conifer-herbaceous classes and decreases
in woody deciduous, mixed woody deciduous-herbaceous, and conifer habitats. The most variance in the
use and cover classes was explained by elevation and
vegetative indices derived from the satellite imagery.
Accuracy ranged from 94% at the coarsest level of
detail to 74% at the finest. The expansion of mixed
conifer classes may increase fire risk to rural homes.
Plant and animal species that specialize on woody
deciduous habitat are probably adversely affected by
reduction in this habitat, as well as by the increase in
rural homes.
Forest Ecology, Culture, and Productivity
Northwest chanterelles, the booklet describes our local species, regional forest management issues, recent
studies, and future research and monitoring needed to
sustain this prized resource.
Phillips, N, BJ Bond, NG McDowell, MG Ryan, and
A Schauer. 2003. Leaf area compounds height-related hydraulic costs of water transport in Oregon
white oak trees. Functional Ecology 17: 832–840.
For tree physiologists and silviculturists. In contrast to
most tree species, the leaf to sapwood area ratio in
Oregon white oak (Quercus garryana) increases with
tree size. The authors investigated whether leaf-specific
hydraulic conductance, crown water flux per leaf area,
and carbon isotope discrimination were reduced in
larger white oak trees with greater leaf area. Water flux
and carbon isotopic discrimination were lower in 25-m
trees than in 10-m trees. The results extend and support the hypothesis that tree-water flux is hydraulically
limited and show that these limitations may not be
reflected by reduced crown leaf area in larger trees.
Poage, NJ, and JC Tappeiner. 2002. Long-term patterns of diameter and basal area growth of oldgrowth Douglas-fir trees in western Oregon. Canadian Journal of Forest Research 32: 1232–1243.
For silviculturists and forest scientists. This study of
505 recently cut old-growth trees at 28 locations in
western Oregon indicates that rapid early and sustained growth of old Douglas-fir trees is extremely
important to the attainment of large diameters at
100–300 years. Diameters of trees at the cut age
had a strong, positive linear relationship to their
diameters and basal area growth rates at age 50 years.
The average periodic basal area increments were
significantly greater in the large trees when young
and remained higher than those of the small trees.
Site factors and establishment class had little influence on tree growth, supporting the hypothesis that
large-diameter old-growth Douglas-fir developed at
low stand densities.
Phillips, NG, MG Ryan, BJ Bond, NG McDowell,
TM Hinckley, and J Cermák. 2003. Reliance on
stored water increases with tree size in three species in the Pacific Northwest. Tree Physiology 23:
237–245.
For tree physiologists. The amount of water used in
daily transpiration that comes from storage (“capacitance”) versus soil was measured in small and large
trees of three species (Douglas-fir, Oregon white
oak, and ponderosa pine). In all species, more water
was stored and used in the larger trees. Utilization
of stored water may help offset hydraulic limitations
resulting from large size. The authors conclude that
water storage is significant in the water and carbon
economy of tall trees and old forests.
Pruyn, ML, ME Harmon, and BL Gartner. 2003.
Stem respiratory potential in six softwood and
four hardwood tree species in the central Cascades
of Oregon. Oecologia 137: 10–21.
For tree physiologists and forest scientists. Tissue-level
respiration (respiratory potential) of mature and oldgrowth trees of 10 species having different sapwood
thickness was measured by an increment-core-based
method under controlled temperature in the laboratory. Measurements of bark (dead and live), sapwood,
and heartwood thickness were used to predict sapwood
volume from stem diameter for four of the species.
Respiratory potential was scaled to the main-bole
level by using these predictions. Core measurement
showed that sapwood respiratory potentials in the
lower bole were 50% higher in species with narrow
sapwood than in those with wide sapwood. This was
not the case for inner bark or within-crown sapwood
respiratory potential. Species with about 20% of the
main bole alive had potential respiration up to 3X
greater than species having >40% alive. The authors
Pilz, D, L Norvell, E Danell, and R Molina. 2003.
Ecology and Management of Commercially Harvested Chanterelle Mushrooms. General Technical
Report PNW-GTR-576, USDA Forest Service,
Pacific Northwest Research Station, Portland OR.
For forest managers, policymakers, mushroom harvesters,
mushroom enthusiasts, and research mycologists. This
synthesis publication discusses chanterelle genera
and species around the world, their international
markets, our current understanding of the organisms,
reasons for declining production in parts of Europe,
and efforts to cultivate them. Focusing on Pacific
32
Forest Ecology, Culture, and Productivity
suggest that the live bole is less metabolically active in
species having large volumes of sapwood.
blocks in sub-boreal forests of Central British Columbia. Blocks were from 0 to 10 years post-harvest.
Instantaneous belowground CO2 flux was lowest in
spring and highest in summer. Cumulative flux was
highest in the 2-year-old cut block, lowest in the
0-year-old cut block, and positively correlated with
soil temperature and biomass present. A few sites had
a significant relationship between soil moisture and
instantaneous belowground CO2 fluxes. Mean soil
temperature was not a good predictor of cumulative
belowground CO2 flux between sites. Plant biomass,
rather than soil temperature, soil moisture, or cut
block age, appears to be the principal factor affecting
belowground CO2 flux in the study.
Puettmann, KJ, and AW D’Amato. 2002. Selecting
plot sizes when quantifying growing conditions in
understories. Northern Journal of Applied Forestry
19: 137–140.
For silviculturists, forest researchers, and ecologists. Plot
sizes are important when the influence of overstory
trees on growing conditions of understory seedlings
is measured. Using different plot sizes, the authors
measured overstory cover (using a “cone” approach) and
basal area at four sites and related these measures to
plot size. Diffuse noninterceptance had a greater range
of values when the cone opening was smaller. Cone
data and 2-year seedling height were related linearly.
This correlation showed that the best plot size differs,
depending on the site, although stand characteristics do
not obviously explain these differences.
Randerson, JT, FS Chapin, JW Harden, JC Neff, and
ME Harmon. 2002. Net ecosystem production: A
comprehensive measure of net carbon accumulation by ecosystems. Ecological Applications 12:
937–947.
For ecologists, biogeochemists, and planetary scientists. Net
ecosystem production (NEP) is generally defined as the
net carbon accumulation by ecosystems. The authors
agree that all the carbon fluxes from an ecosystem, not
just the balance between ecosystem respiration and
gross primary production, must be considered in calculating NEP. Net biome productivity (NBP) is the same
as NEP at regional or global scales. NEP estimates
made at many scales can be compared directly by using
the authors’ conceptual framework.
Pypker, TG, and AL Fredeen. 2002. Ecosystem CO2
flux over two growing seasons for a sub-boreal
clearcut 5 and 6 years after harvest. Agricultural
and Forest Meteorology 114: 15–30.
For ecologists and forest scientists. Two approaches, a
Bowen ratio energy balance method and a component
flux model, were used to measure ecosystem-level
growing season CO2 fluxes in a 6-year-old vegetated
sub-boreal clearcut. Measurements were taken in
2000 from May 24 to Sep 20 and compared to
measurements from 1999 from June 27 to Sep 3.
Results indicated the clearcut was a source of CO2 in
2000 and a sink in 1999. The clearcut had photosynthetic uptake was about equal in both years, but
soil surface CO2 efflux was higher in 2000. This
young forest appeared to be a net source of CO2.
Rich, JJ, RS Heichen, PJ Bottomley, K Cromack, Jr,
and DD Myrold. 2003. Community composition
and functioning of denitrifying bacteria from adjacent meadow and forest soils. Applied and Environmental Microbiology 69: 5974–5982.
Pypker, TG, and AL Fredeen. 2003. Belowground
CO2 efflux from cut blocks of varying ages in
sub-boreal British Columbia. Forest Ecology and
Management 172: 249–259.
For forest managers, forest scientists, and ecologists.
From May to October 2000, instantaneous measures
of belowground CO2, temperature, and moisture
were taken in a mature stand and seven vegetated cut
33
For soil ecologists, microbiologists, and community ecologists. Measuring denitrifying enzyme activity as a proxy
for function and using nosZ as a marker for denitrifying bacteria, the authors examined spatial gradients
and community composition, their relation to ecological properties, and phylogenetic relationships among
denitrifiers in adjacent meadow and forest soils. The
meadow and forest soils exhibited marked differences
in functional and structural communities. There were
47 unique denitrifying genotypes. Community com-
Forest Ecology, Culture, and Productivity
position was influenced by process rates and vegetation
types. Both composition of the denitrifying community
and environmental factors may contribute to variability of denitrification rates.
Ripple, WJ, and RL Beschta. 2003. Wolf reintroduction, predation risk, and cottonwood recovery in
Yellowstone National Park. Forest Ecology and
Management 184: 299–313.
crops on a scientific basis so that GEOs can be used
to improve sustainability in agricultural systems.
Rothe, A, K Cromack, SC Resh, E Makineci, and Y
Son. 2002. Soil carbon and nitrogen changes
under Douglas-fir with and without red alder. Soil
Science Society of America Journal 66: 1988–
1995.
For ecologists, forest scientists, and silviculturists. To
investigate the influence of red alder (RA) on carbon
(C) and nitrogen (N) pools, the authors sampled three
plots consisting of only Douglas-fir (DF) and three
plots with a 25% RA/75% DF mix in 1980 and
1999. The experimental design allowed both homeseries and spatial-series analysis. The N pool grew
with an average accretion of 8.7 g/m2/year. Resin N
mineralization in mixed plots was 10 times greater
than in the pure DF plots. Foliar N concentration
of DF was not influenced. In order to detect changes
in the soil N pool, plot pairing was necessary and
only large effects could be determined with statistical
significance. Total soil N accretion in mixed plots was
28% greater than the minimum detectable difference
after 19 years. Carbon differences, however, would
not be detectable for another 30 years.
For forest ecologists, forest managers, and wildlife and
conservation biologists. Wolves were reintroduced into
Yellowstone National Park (YNP) in the winter of
1995–1996 after an absence of about 70 years.
The authors examined the potential influence of
interaction between wolves and elk on growth of
young riparian cottonwoods (Populus spp.) and associated woody plants in YNP. Comparing heights
of woody plants in photographs taken before 1998
with heights in 2001–2002 photos, they found
that height increased on six of the eight sites in the
study area. They measured browsing intensity and
cottonwood height on sites with relatively high and
relatively low risk of predation by wolves. In general,
browsing intensities were lower and plants were taller
on sites where elk were at high risk of predation than
they were where predation risk was relatively low. The
study proves rare empirical evidence of the indirect
effects of a top carnivore in a terrestrial food chain
that supports theories on predation risk effects and
top-down control. Reintroduction of wolves into
YNP may help ensure the restoration of riparian species and preservation of biodiversity.
Rothe, A, J Ewald, and DE Hibbs. 2003. Do admixed
broadleaves improve foliar nutrient status of conifer tree crops? Forest Ecology and Management
172: 327–338.
For silviculturists, forest scientists, and ecologists. The
authors looked at eight broadleaf and conifer mixed
stands in Germany and Oregon to see if the nutritional status of the conifers would improve among
the broadleaf trees. Needles from 20–30 trees at
each site and needle mass, N, P, K, Ca, and Mg were
analyzed. Measurements were compared based on the
basal area of the deciduous trees within an 8-m circle
of the measured tree. The conifers showed no improvement in foliar nutrition as a function of proximity to broadleaves, which agreed with other studies.
Broadleaf admixtures apparently provide little nutritional benefit to admixed conifer.
Ronald, P, and SH Strauss. 2003. Moving the debate
on genetically engineered crops forward. American
Society of Plant Biologists Newsletter 30(3): pp. n.a.
For plant biologists, molecular geneticists, and agricultural technologists. The development of genetically
engineered organisms (GEOs) in crop species has
turned plant breeding into a very controversial area.
Responding to an earlier Perspectives column in the
Newsletter, the authors present arguments against
seven “scientific myths” that they feel are impeding
discussion and progress in research on and use of
genetically engineered crops. They call on the Society
to encourage evaluation of GEO crops or classes of
Schowalter, TD, and LM Ganio. 2003. Diel, seasonal
and disturbance-induced variation in invertebrate
34
Forest Ecology, Culture, and Productivity
assemblages, pp. 315–328 in Arthropods of Tropical Forests: Spatio-temporal Dynamics and Resource
Use in the Canopy, Y Basset, V Novotny, SE
Miller, and RL Kitching, eds. Cambridge University Press, Cambridge, UK.
For invertebrate ecologists. The authors sampled canopy
invertebrate assemblages and leaf area missing in
representative early and late successional tree species. At the Luquillo Experimental Forest Long Term
Ecological Research Programme site in Puerto Rico,
they sampled during day and night, in wet and dry
seasons, and in plots that had incurred light or severe
hurricane damage. At the Fort Sherman canopy crane
in Panama, they sampled during wet and dry seasons. Many species and functional groups differed in
abundance among tree species, as expected, and leaf
area missing varied significantly both seasonally and
annually. Unexpectedly, diel and seasonal patterns of
abundance did not show differences in any species.
Predators and detritivores were more abundant during drought, and predators also increased after hurricanes. The authors discuss interaction among tree
species, years, and various invertebrate species and
functional groups.
Schowalter, TD, YL Zhang, and JJ Rykken. 2003.
Litter invertebrate responses to variable density
thinning in western Washington forest. Ecological
Applications 13: 1204–1211.
For ecologists, silviculturists, and entomologists. Seven
and eight years after variable density thinning of the
Douglas-fir forest overstory, pitfall traps were placed
in four thinned and four unthinned units in each of
two sites having different management histories. Peak
abundance for most species was in summer. Initial
structure of the thinning treatment affected the way
in which the invertebrates were affected. Only 2 of
29 taxa responded significantly to thinning. Three
spider species were identified as potential indicators of thinning treatment. Species assemblages in
treated units were significantly different from those
in control units and from each other, but combined
taxa and function group did not so differ. Previous
management practices strongly influenced the effect
of variable density thinning on litter invertebrates.
Schwarz, PA, TJ Fahey, and CE McCulloch. 2003.
Factors controlling spatial variation of tree species
abundance in a forested landscape. Ecology 84:
1862–1878.
For forest ecologists. The authors investigated whether
biological factors, such as disturbance and competition, within a local neighborhood (neighborhood factors) might influence spatial patterns of tree species
in forests. They measured tree abundance patterns
and environmental and disturbance factors expected
to regulate the patterns in the Hubbard Brook Experimental Forest in New Hampshire. Their models
of environmental and disturbance effects explained
26–62% of variation in abundance among the
seven tree species that contribute 90% of the total
basal area in the Forest. The abundance patterns of
six of the seven species suggested that factors other
than environment and disturbance influence spatial
patterns. The data were consistent with regulation
by seed dispersal distance and root sprouting. The
authors argue that neighborhood factors significantly
influence patterns of tree species in forests.
Shatford, J, D Hibbs, and L Norris. 2003. Identifying
plant communities resistant to conifer establishment along utility rights-of-way in Washington
and Oregon, U.S. Journal of Arboriculture 29(3):
172–176.
For arboriculturists and utility planners. The authors
sampled vegetation in 1376 plots in electrical rights
of way at three sites to determine percent cover of all
woody species and density of tree stems. Douglas-fir
seedlings were not randomly distributed among community types. They also occurred less frequently in
communities occupied by dense vegetation, including
some low-growing communities, such as are desirable
under utility lines.
Skinner, JS, R Meilan, CP Ma, and SH Strauss. 2003.
The Populus PTD promoter imparts floral-predominant expression and enables high levels of
floral-organ ablation in Populus, Nicotiana and
Arabidopsis. Molecular Breeding 12: 119–132.
35
For molecular geneticists, biotechnologists, tree geneticists,
and plant breeders. The Populus PTD gene was tested
Forest Ecology, Culture, and Productivity
for its ability to genetically engineer reproductive
sterility. Using the promoter resulted in sterile plants
with otherwise normal growth at high frequency in
Arabidopsis, tobacco, and poplar. Test of biomass
production in greenhouse-grown tobacco confirmed
strong oral specificity of the promoter. Using this
gene may allow transgene confinement without decreasing yield.
action. Unexpectedly, emergent behaviors were found
at all levels, rather than ending as one moved from
the patch to the landscape level. Their magnitude depended on the level of spatial interaction and the type
and intensity of the processes included in the simulation. They conclude that using an additive approach
may not capture some C dynamics at broad scales in
fragmented landscapes, although spatial interactions
in some cases may be small enough to ignore.
Smithwick, EAH, ME Harmon, SM Remillard, SA
Acker, and JF Franklin. 2002. Potential upper
bounds of carbon stores in forests of the Pacific
Northwest. Ecological Applications 12: 1303–
1317.
For forest researchers and forest ecologists. The authors
put an upper bound (or limit) on carbon (C) storage
in the Pacific Northwest (PNW) of the United States
using field data from old-growth forests, which are
near steady-state conditions. The upper bounds were
higher than current estimates of C stores, presumably
because of both natural and anthropogenic disturbances. This finding indicates a potentially substantial and economically significant role of C sequestration in the region. Coastal Oregon stands stored
more C, on average, than stands in eastern Oregon
and coastal Washington. Similarly, Oregon Cascades
stands stored more, on average, than Washington
Cascades stands. A maximum of 338 Mg C/ha (total
ecosystem C­100) could be stored in PNW forests in
addition to current stores.
Smithwick, EAH, ME Harmon, and JB Domingo.
2003. Modeling multiscale effects of light limitations and edge-induced mortality on carbon
stores in forest landscapes. Landscape Ecology 18:
701–721.
For landscape ecologists and those interested in carbon
cycles. Usually spatial interactions are not considered
in analysis of carbon (C) dynamics at broad scales, the
assumption being that C dynamics can be modeled
within homogenous patches and then added to predict
dynamics on a broad scale. The authors used a forest
process model, light and wind processes, and artificial
forest landscapes with spatial structures ranging from
the patch to the landscape to check for possible emergent behaviors resulting from pattern-process inter-
Spears, JDH, SM Holub, ME Harmon, and K Lathja.
2003. The influence of decomposing logs on soil
biology and nutrient cycling in an old-growth
mixed coniferous forest in Oregon, U.S.A. Canadian Journal of Forest Research 33: 2193–2201.
For soil biologists, ecologists, silviculturists, and those interested in nutrient cycling. The authors compared nutrients in leachates of coarse woody debris (CWD) with
leachates from the forest floor in the central Cascade
Range in Oregon. They also compared CWD leachates over a decay chronosequence and among CWD
of four species. They found few differences among
CWD and forest floor leachates or among species
and found no differences in nitrogen, phosphorus,
microbial biomass, Biolog plate assays, or enzyme
activity in soils. Differences were largest during the
middle decay classes. Their results suggest that either
CWD has no long-term effect and does not contribute much organic matter to the soil or its effect is so
prolonged that no spatial effect can be detected.
Spies, TA, DE Hibbs, JL Ohmann, GH Reeves, RJ Pabst,
FJ Swanson, C Whitlock, JA Jones, BC Wemple,
LA Parendes, and BA Schrader. 2002. The ecological basis of forest ecosystem management in
the Oregon Coast Range, pp. 31–67 in Forest and
Stream Management in the Oregon Coast Range,
SD Hobbs, JP Hayes, RL Johnson, GH Reeves, TA
Spies, JC Tappeiner II, and GE Wells, eds. Oregon
State University Press, Corvallis.
36
For forest ecologists and forestland managers. The authors discuss twelve major ecological themes that they
believe underlie ecologically based forest management in the Oregon Coast Range. The themes are
grouped into three categories: ecosystem patterns and
history, disturbance and vegetation development, and
Forest Ecology, Culture, and Productivity
landscape interactions. In concluding, they discuss
how understanding natural processes can contribute
to reaching ecosystem goals. Their aim is to help define the ecological constraints on the output of social
and economic values in the Oregon Coast Range.
should be highly restricted or forbidden and some
that are quite safe. Breeders are important in providing species variation, and genetic engineers must work
extremely hard to move agronomic traits further than
breeders can. Tree breeders know that natural selection provides a safeguard for undesirable traits. Gene
flow matters, but the challenge is to decide how little
gene flow is little enough.
Strauss, SH. 2003. Genetic technologies: Genomics,
genetic engineering, and domestication of crops.
Science 300: 61–62.
For plant geneticists, forest scientists, policy makers,
and researchers. Genomics allows use of recombinant DNA techniques to domesticate crops in ways
similar to conventional breeding. Genomics-guided
transgenes will be based increasingly on nature or
homologous genes from related species, rather than
on exogenous genes. The future seems to hold great
promise for genomics-guided transgene modifications
in agriculture. The process needs to be tried in the
field, but regulations are becoming increasingly strict
in many parts of the world. This impedes the delivery of benefits from the genomics revolution to the
public. The author proposes some ways to facilitate
field testing.
Strauss, SH. 2003. Democratization is more than
lower prices–Response. Science 301: 167–167.
Strauss, SH. 2003. Risks of genetically engineered
crops–Response. Science 301: 1846–1847.
For scientists and policymakers interested in genetic
engineering of crops. The author responds to two letters
written to Science commenting on his contribution
to the journal’s Policy Forum (Science 300: 61–62)
about regulation of genetically engineered crops.
Stuart-Smith, AK, and JP Hayes. 2003. Influence of
residual tree density on nest predation of artificial
and natural songbird nests. Forest Ecology and
Management 183: 159–176.
For forest managers and ecologists. Nest success and
nest predation were measured in natural and artificial nests in burned and unburned stands varying in
residual tree density in British Columbia. In one year
of the 2-year study, predation increased slightly in the
logged stands. There was no increased predation in
the burned stands. Overall, nest predation was not related to residual tree density, suggesting that management practices that retain trees after timber harvest
do not create “ecological traps” for songbirds.
Styles, JM, J Lloyd, D Zolotoukhine, KA Lawton, N
Tchebakova, RJ Francey, A Arneth, D Salamakho,
O Kolle, and E-D Schulze. 2002. Estimates of regional surface carbon dioxide exchange and carbon
and oxygen isotope discrimination during photosynthesis from concentration profiles in the atmospheric boundary layer. Tellus B 54: 768–783.
For researchers in boundary layer meteorology and plantatmosphere exchange. An atmospheric boundary layer
budget was used to infer regional surface fluxes
of CO2 and isotopic discrimination from aircraft
concentration profiles. The two methods agreed
within 10%. When corrected for air loss out of
the integrating column, the methods agreed within
35% of each other. 13C values for discrimination
complied with the expected range, but 18O values for
discrimination varied considerably.
Strauss, SH. 2003. Regulating biotechnology as
though gene function mattered. BioScience 53:
453–454.
For geneticists, genetic engineers, and ecologists. Evaluation of genetically engineered crops on a per-crop
basis is costly to society. The author discusses an
article by Jim Hancock in the same issue of BioScience, proposing that some crops should be exempted
from requirements for most kinds of environmental
studies. Hancock sorts transgenic crops into biologically rational groups, recognizing some genes that
Styles, JM, MR Raupach, GD Farquhar, O Kolle, KA
Lawton, WA Brand, RA Werner, A Jordan, E-D
Schulze, O Shibistova, and J Lloyd. 2002. Soil
and canopy CO2, 13CO2, H2O and sensible heat
37
Forest Ecology, Culture, and Productivity
For forest scientists and ecologists. The authors determined soil respiration and shifts in microbial
biomass in light fractions (LF), heavy fractions
(HF), whole soils (WS), and physically recombined
light and heavy fractions (RF) of mineral soils from
forests in Washington and Oregon. The summed
fraction (SF) was calculated from the incubation
results of LF and HF. Carbon (C) concentration
had the pattern LF>RF>HF. Cumulative respiration of physical fractions showed the same pattern
when considered per gram of substrate, but LF
values per gram of initial C were not different from
those of HF. Recalcitrance of HF may be similar to
that of LF. Respiration of the SF was not different
from that of the WS, whether expressed per gram of
substrate or per gram of initial C. The HF accounted for 35% of the total respiration within the SF.
Lower respiration in the RF, compared with the WS
and the SF, may be due to an antagonistic interaction between different microbial communities that
degrade LF and HF and alteration of normal spatial
relations of the fractions in the laboratory. The
density-separation technique seems to isolate soil
organic matter fractions effectively.
flux partitions in a forest canopy inferred from
concentration measurements. Tellus B 54: 655–676.
For researchers in micrometeorology and plant-atmosphere
exchange. A multi-layer canopy model coupled to a
Lagrangian transport model was applied to a coniferous forest in Siberia. Concentration profiles within
the canopy were assimilated and parameters within
the canopy model were optimized in order to infer the
source distribution within the canopy of CO2, water
vapor, heat, and carbon and oxygen isotopes of CO2.
Low temperatures that occur during snow melt caused
depressed stomatal conductance and maximum
photosynthetic capacity. Radiation penetrated further
because of leaf clumping and penumbra, contrary to
theoretical predictions. Important stability effects
were seen in the morning and the evening. Limits
to the inversion were little vertical structure in the
concentration profiles and codependence of canopy
parameters.
Suzuki, N, and JP Hayes. 2003. Effects of thinning on
small mammals in Oregon coastal forests. Journal
of Wildlife Management 67(2): 352–371.
For ecologists and forest managers. The authors looked
at the poorly understood effects of thinning densely
stocked Douglas-fir forests of the Pacific Northwest
on forest-floor small mammals. Two pitfall-trapping
studies were conducted in the Oregon Coast Range.
The experimental study, which dealt with stands for
the first 2 years after thinning, found captures increased for 4 of 12 species examined and decreased
for 1. The retrospective study, which dealt with
previously thinned (7–24 years before the study)
and unthinned stands, looked at 9 species and found
captures increased for 5 and decreased for none in
thinned stands, compared with unthinned stands.
More small mammals were captured in total in the
previously thinned stands. Thinning had no significant negative effects on the species examined, and
several species were positively affected by thinning.
Tappeiner, JC, II, WH Emmingham, and DE Hibbs.
Silviculture of Oregon of Oregon Coast Range
Forests, pp. 172–190 in Forest and Stream Management in the Oregon Coast Range, SD Hobbs,
JP Hayes, RL Johnson, GH Reeves, TA Spies, JC
Tappeiner II, and GE Wells, eds. Oregon State
University Press, Corvallis.
For silviculturists and forestland managers. Silviculture
includes reforestation after disturbances and thinning
stands in order to increase diversity or vigor, with the
objectives of producing wood, growing large trees, and
enhancing habitat for plant and animals, often simultaneously. Putting silvicultural principles into practice
requires much site-specific evaluation. The authors
review the biological and ecological basis of practical
silviculture in the Oregon Coast Range and discuss
how silviculture can be used to ensure regeneration
after major disturbances, to manage forest for wood
products and other values, to develop structural complexity, and to manage riparian zones. They also treat
the silviculture of alder and cottonwood plantations.
Swanston, CW, BA Caldwell, PS Homann, L Ganio,
and P Sollins. 2002. Carbon dynamics during a
long-term incubation of separate and recombined
density fractions from seven forest soils. Soil Biology and Biochemistry 34: 1121–1130.
38
Forest Ecology, Culture, and Productivity
ic CO2 on NEE are usually limited by N availability.
Rates of carbon sequestration over the past 200
years were a function of the rate of change in CO2
concentration for old sites with low rates of N-dep.
The model was effective at estimating between-site
variation in leaf area index, but not as effective with
variation of evapotranspiration between- and withinsite. The model does have some biases.
Tchebakova, NM, O Kolle, D Zolotoukhine, A Arneth, JM Styles, NN Vygodskaya, E-D Schulze, O
Shibistova, and J Lloyd. 2002. Inter-annual and
seasonal variations of energy and water vapour
fluxes above a Pinus sylvestris forest in the Siberian
middle taiga. Tellus B 54: 537–551.
For researchers in micrometeorology and plant-atmosphere
exchange. Seasonal and annual fluxes of water vapour
and energy in a Siberian pine forest were assessed
from 1998 to 2000. Surface energy exchange
characteristics were distinctly seasonal. In the early
spring, 80% of the net radiation was partitioned
for sensible heat. When photosynthesis began,
evaporation rates increased and Bowen ratios (β)
decreased. During the summer, sensible heat fluxes
usually exceeded latent heat fluxes, keeping β above
2. Daily evaporation was, on average, 1.25 mm/d, and
precipitation was 230 mm for the growing period.
Only about 35% of the equilibrium evaporation rate
was represented. Usually there was a positive hydrological balance of 40 mm, but evaporation always
exceeded precipitation by 20–40 mm in at least one
calendar month during the summer. Growing season
surface conductances of dry or cool months varied
between 0.15 and 0.20 mol/m2s and 0.30 and 0.35
mol/m2s in moist and warm months.
Thornton, PE, BE Law, HL Gholz, KL Clark, E Falge,
DS Ellsworth, AH Golstein, RK Monson, D
Hollinger, M Falk, J Chen, and JP Sparks. 2002.
Modeling and measuring the effects of disturbance
history and climate on carbon and water budgets
in evergreen needleleaf forests. Agricultural and
Forest Meteorology 113: 185–222.
For ecologists and forest scientists. The authors evaluated the effects of disturbance history, climate,
and changes in atmospheric carbon dioxide (CO2)
concentration and nitrogen deposition (N-dep) on
carbon and water fluxes in seven North American evergreen forests. As modeled by the ecosystem process
model Biome-BGC, net ecosystem carbon exchange
(NEE) is mostly a function of disturbance history and
is also affected by site climate, vegetation ecophysiology, and changing atmospheric CO2 and N-dep.
Fluxes after a disturbance can vary greatly in size and
timing. The modeled effects of increasing atmospher-
Trappe, JM, L Dominguez, M Castellano, E Cazares, T
Lebel, and AW Claridge. 2002. Hypogeous fungi and
the Gondwanan connection. Inoculum 53(3): 56.
For researchers in forest fungus taxonomy and ecology.
Field and laboratory research on truffle-like fungi in
the Southern Hemisphere shows strong relationships
between families and genera of Australia, New Zealand, and South America. The authors hypothesize
that many of these fungi evolved when those land
masses were part of the supercontinent Gondwana,
before separation by continental drift.
Trappe, JM, and AW Claridge. 2003. Australasian
sequestrate (truffle-like) fungi. 15. New species
from tree line in the Australian Alps. Australasian
Mycologist 22: 27–38.
For mycologists. Eighteen species of hypogeous fungi
were found at tree lines in the Australian Alps. Six of
these were new and are described in this paper. They
are all probably in ectomycorrhizal association with
Eucalytpus niphophila, the dominant tree species in the
area. Five of the six are known from lower elevations.
Treuhaft, RN, GP Asner, BE Law, and S Van Tuyl.
2002. Forest leaf area density profiles from the
quantitative fusion of radar and hyperspectral
data. Journal of Geophysical Research—Atmospheres 107: art. no. 4568.
39
For ecosystem modelers and ecologists. The authors
combined data from NASA Airborne Synthetic
Aperture Radar (AIRSAR) and NASA Airborne
Visible and Infrared Imaging Spectrometer (AVIRIS)
for three forest plots in Central Oregon. Leaf area
density (LAD) was extracted from the radar and
hyperspectral data that were gathered. LAD data were
represented as a function of height. The remotely
Forest Ecology, Culture, and Productivity
sensed data were compared with field measurements
and checked for accuracy. The two types of data were
usually within 0.02 m2/m3 and always within the 1–2
standard error range. With further development in
processes and models, LAD effectiveness will improve.
Treuhaft, RN, GP Asner, and BE Law. 2003. Structure-based forest biomass from fusion of radar and
hyperspectral observations. Geophysical Research
Letters 30: art. no. 1472.
For ecologists and forest scientists. Remotely sensed
profiles of canopy leaf area density were used to estimate forest biomass of 11 structurally diverse stands
in Central Oregon. The field and remotely sensed
measurements of biomass were in agreement, yielding
a level of 25 tons/ha, 16% of the average stand biomass, which was significant with >99.5% confidence.
This opens up the possibility of a set of model functions that will enable global, structure-based biomass
remote sensing.
duction (NEP) at the landscape scale, using remote
sensing, ground measurements, and ecosystem modeling. Most areas in the study were carbon sinks, with
an average annual NEP for the study area of 230 g
C/m2. The NEP was strongly positive because harvesting in the area had been reduced over a decade.
Interannually, NEP varied by a factor or two, 38%
less than for a single point.
Turner, DP, WD Ritts, WB Cohen, ST Gower, MS
Zhao, SW Running, SC Wofsy, S Urbanski, AL
Dunn, and JW Munger. 2003. Scaling gross
primary production (GPP) over boreal and deciduous forest landscapes in support of MODIS GPP
product validation. Remote Sensing of Environment
88: 256–270.
For landscape ecologists, silviculturists, and those interested
in remote sensing. The 2001 gross primary production
(GPP) product of the Moderate Resolution Imaging Radiometer (used by the NASA Earth Observing
System to monitor seasonality of terrestrial vegetation)
was compared with scaled GPP estimates from ground
measurements at two forested sites. The GPP phenology from MODIS started earlier at the hardwood forest
site than the scaled GPP indicated, and the MODIS
summertime GPP was generally lower than the scaled
values. The phenologies generally agreed at the boreal
forest site, but the midsummer MODIS GPP was generally higher than the ground-based values. Groundbased GPP scaling could improve the parameterization
of light-use efficiency in satellite-based algorithms for
monitoring GPP.
Turner DP, ST Gower, M Gregory, and TK Maiersperger. 2002. Effects of spatial variability in light
use efficiency on satellite-based NPP monitoring.
Remote Sensing of Environment 80: 397–405.
For ecologists and remote sensing scientists. Light use
efficiency (LUE) algorithms, which potentially can be
used to monitor global net primary production (NPP)
using satellite-borne sensors, may introduce error
because they are applied at coarse spatial resolution
that may subsume significant heterogeneity in vegetation LUE. The authors examine the effects of spatial
heterogeneity on an LUE algorithm over a 25-km2
area of corn and soybean. Their results suggest several
approaches to account for land cover heterogeneity when LUE algorithms are implemented at coarse
resolution.
Turner, DP, S Urbanski, D Bremer, SC Wofsy, T Meyers, ST Gower, and M Gregory. 2003. A crossbiome comparison of light use efficiency for gross
primary production. Global Change Biology 9:
383–395.
For ecologists. The authors describe differences among
four biomes in maximum light use efficiency (LUE)
and in environmental controls on LUE. All the
sites had a hyperbolic relationship of gross primary
production (GPP) and absorbed photosynthetically
active radiation (APAR) except the tallgrass prairie,
which was linear. The agricultural field had higher
values of light use efficiency (eg ) than the boreal for-
Turner, DP, M Guzy, MA Lefsky, S VanTuyl, O Sun,
C Daly, and BE Law. 2003. Effects of land use
and fine scale environmental heterogeneity on net
ecosystem production over a temperate coniferous
forest landscape. Tellus B 55: 657–668.
For ecologists and carbon cycle modelers. The authors
present an approach to mapping net ecosystem pro-
40
Forest Ecology, Culture, and Productivity
est. Higher APAR was associated with lower eg. Maximum daily temperature and vapor pressure deficit did
not correlate well with eg. Decline in eg was detected
at the end of the growing season at the agricultural
site and in August during a soil drought at the tallgrass prairie.
Turner, MG, SL Collins, AE Lugo, JJ Magnuson, TS
Rupp, and FJ Swanson. 2003. Disturbance dynamics and ecological response: The contribution of
long-term ecological research. Bioscience 53: 46–56.
For ecologists and those interested in disturbance dynamics.
To illustrate the importance of long-term ecological
research in understanding disturbance dynamics, the
authors present three case studies from Long Term
Ecological Research (LTER) network sites: a temperate
lake invaded by exotic species, a tropical forest hit by a
hurricane, and a fire-influenced temperate grassland.
The LTER network encompasses ecosystems affected
by a wide range of disturbances, allows measurement of
ecosystem disturbances against a long-term baseline,
allows slow or rare events to be observed, facilitates
multiple research approaches, provides a focus for
modeling disturbance dynamics, and contributes to
management of resources.
Venterea, RT, GM Lovett, PM Groffman, and PA
Schwarz. 2003. Landscape patterns of net nitrification in a northern hardwood-conifer forest. Soil
Science Society of America Journal 67: 527–539.
additional 10% of variance. Wide variation in NO3results from a combination of multiple biotic and
abiotic factors. The study provides a quantitative
basis for estimates of how changes in vegetation and
climate may influence N cycling.
Vesely, DG, and WC McComb. 2002. Salamander
abundance and amphibian species richness in
riparian buffer strips in the Oregon Coast Range.
Forest Science 48: 291–297.
For ecologists, zoologists, and forest scientists. The
authors researched total salamander abundance,
amphibian species richness, and sampling proportions
for five salamander species in 17 managed stands
and 12 unlogged forests near streams to determine
the effectiveness of riparian buffer strips. The results
showed that four of the five species, along with total
salamander abundance and amphibian species richness, were affected by forest practices. Riparian buffer
strips are helpful for several salamander species, but
state forest practices regulations do not currently
provide for riparian zones that are wide enough to
prevent local declines in the diversity of amphibian
communities.
Wagai, R, and P Sollins. 2002. Biodegradation and
regeneration of water-soluble carbon in a forest
soil: Leaching column study. Biology and Fertility
of Soils 35: 18–26.
For soil scientists and soil microbiologists. The authors
hypothesized that water-soluble carbon (C) provides
a major substrate for soil microbes. The dissolved
organic C was constant in leachates collected every 2
weeks over 20 weeks. The pool of water-extractable
carbon declined by 31–40% over the 20 weeks The
presence of tree seedlings decreased the amount and
biodegradability of leachates and extracts. Leachable
C did not appear to provide significant substrate for
heterotrophic soil respiration in the system, but the
role of water-extractable C as a substrate was unclear.
F
or ecosystem modelers and forest scientists. Net
nitrification rates (NR) and nitrate (NO3- ) concentrations were taken at 100 plots across a 3160-ha
hardwood-conifer forest in central New Hampshire.
Their associations with physiographic features and
vegetation abundances were investigated. Both measures varied by a factor of 150. NO3- production was
greater at higher plot elevations or more southerly
aspects, with more sugar maple and striped maple, and
with fewer conifers. These factors explained 52% of
the variance in NR in regression models. Landscape
patterns were affected by higher soil water contents,
nitrogen (N) mineralization rates, total N concentrations in higher elevation plots, and higher N mineralization and respiration rates in more south-facing
plots. Soil C/N ratios and landscape explained an
Waldien, DL, JP Hayes, and BE Wright. 2003. Use of
conifer stumps in clearcuts by bats and other vertebrates. Northwest Science 77(1): 64–71.
41
For forest managers, vertebrate biologists, and ecologists.
Use of conifer stumps by vertebrates was monitored
Forest Ecology, Culture, and Productivity
in four stands in the Oregon Cascades. Vertebrate use
of stumps was low. The highest use was in relatively
large Douglas-fir stumps located in open areas. Ten
species of vertebrates used bark crevices. Bats using
stumps were primarily long-eared myotis. Maintaining present and future availability of snags should be
the focus of management of bat roost structures.
scales. The patterns were derived from 10 tree-ringbased fire history studies located west of the crest of
the Cascade Range in the Pacific Northwest. Anthropogenic change, climate, and the degree of stand/fuel
development appear to have interacted in their influence on temporal variation in fire regimes. Patterns
of temporal variation in area burned among the 10
studies reveal the roughly synchronous nature of fire
in the region, which has important implications for
our understanding of landscape dynamics before
settlement.
Waring, RH, and N McDowell. 2002. Use of physiological process model with forestry yield tables to
set limits on annual carbon balances. Tree Physiology 22: 179–188.
Well, R, J Augustin, K Meyer, and DD Myrold. 2003.
Comparison of field and laboratory measurement
of denitrification and N2O production in the saturated zone of hydromorphic soils. Soil Biology and
Biochemistry 35: 783–799.
For tree physiologists and those studying carbon balance.
The authors used a process-based model, yield tables,
and local weather station to set limits on annual
carbon (C) fluxes for forest of different ages. When
constrained to match stand dynamics, the mode can
provide reasonable annual estimates of gross photosynthesis for a specified climate. With additional
assumptions and estimates, maximum net annual
ecosystem exchange can also be predicted.
For researchers in soil nitrogen cycling. The authors
tested a new method for measuring in situ denitrification under field conditions in many water-saturated
subsoils with a range of biochemical properties. The
in situ denitrification rates ranged from 1–2800
μgN/kg/d, compared with 1–1700 μgN/kg/d in the
laboratory. Both the laboratory and field results are
considered valid, and both were ineffective at determining the proportion of N2O in the total denitrification output.
Waring, RH, NC Coops, JL Ohmann, and DA Sarr.
2002. Interpreting woody plant richness from
seasonal ratios of photosynthesis. Ecology 83:
2964–2970.
For ecologists and forest scientists. A satellite-driven
process model provided the data to predict gross photosynthesis and the most constraining environmental
variable and, thereby, explain the wide variation of
species richness in Oregon forests. Species richness
was highest on sites where 60–70% of light was intercepted by vegetation. Evergreen and deciduous forest photosynthesis peaked in the spring and summer.
Seasonal ratio of gross photosynthesis varied from
<1 to >5. Mild, moist spring weather conditions and
summer drought yielded the highest species richness.
Weisberg, PJ, and FJ Swanson. 2003. Regional synchroneity in fire regimes of western Oregon and
Washington, USA. Forest Ecology and Management 172: 17–28.
For forest historians and forest ecologists. Temporal patterns of area burned at 25-year intervals over 600
years were studied to understand the implications of
fire history on forest dynamics over stand to regional
Whitbeck, KL, MA Castellano, JW Spatafora, E
Cazares, and JM Trappe. 2002. Systematics of the
genus Gymnomyces (Russulales, Basidiomycota).
Inoculum 53(3): 59.
For researchers in forest fungus taxonomy and ecology.
Analysis of DNA from species in this genus of truffle-like fungi indicated that different species of Gymnomyces apparently are derived from several different
sources in their ancestral mushroom genus Russula.
In addition, there may be fewer phylogenetic species
than previously thought.
Wilson, K, A Goldstein, E Falge, M Aubinet, D Baldocchi, P Berbigier, C Bernhofer, R Ceulemans,
H Dolman, C Field, A Grelle, A Ibrom, BE Law,
A Kowalski, T Meyers, J Moncrieff, R Monson,
W Oechel, J Tenhunen, R Valentini, and S Verma.
2002. Energy balance closure at FLUXNET
42
Forest Ecology, Culture, and Productivity
sites. Agricultural and Forest Meteorology 113:
223–243.
Wirth, TA, and DA Pyke. 2003. Restoring forbs for
sage grouse habitat: Fire, microsites, and establishment methods. Restoration Ecology 11: 370–377.
For environmental scientists and ecologists. The authors did a comprehensive evaluation of energy
balance closure across 22 sites and 50 site-years in
FLUXNET, using statistical regression of turbulent
energy fluxes and solving for the energy balance
ratio. Most sites lacked closure; the mean imbalance
was about 20%. The imbalance was apparent in all
measured climates and vegetation types. Closure was
better with turbulent intensity (friction velocity) and
worst at night. Either fluxes of sensible and latent
heat were underestimated, available energy was overestimated, or both. While it is not known what error
caused the imbalance, circumstantial evidence indicates a link between the imbalance and CO2 fluxes.
For wildlife biologists and restoration biologists. The
availability of certain preferred and nutritious forb
species may limit sage grouse productivity. The
authors determined the suitability of three species
of forbs for revegetation projects to improve grouse
habitat. They determined emergence, survival, and
reproduction of Crepis modocensis, C. occidentalis,
and Astragalus purshii. They used combinations
of two methods each of establishment (seeding or
transplanting), site preparation (burned or not), and
microsite (mound or interspace) in an Artemisia
tridentata vegetation association in south central
Oregon. Emergence of seeded Astragalus was lower
than that of either seeded Crepis species (each 38%).
Emergence of Crepis was highest in mounds in unburned areas. Seedling survival was higher in burned
areas and in mounds, however, so plant establishment
was better in burned mounds. Fire improved survival
of emergent A. purshii seedlings and transplants of all
three species. Crepis may be a good revegetation option for improving sage grouse habitat in the region,
but Astragalus requires more research.
Wilson, KB, DD Baldocchi, M Aubinet, P Berbigier,
C Bernhofer, H Dolman, E Falge, C Field, A
Goldstein, A Granier, A Grelle, T Halldor, D Hollinger, G Katul, BE Law, A Lindroth, T Meyers, J
Moncrieff, R Monson, W Oechel, J Tenhunen, R
Valentini, S Verma, T Vesala, and S Wofsy. 2002.
Energy partitioning between latent and sensible
heat flux during the warm season at FLUXNET
sites. Water Resources Research 38: art. no. 1294
For forest researchers and ecologists. The Bowen ratio
[beta (B) = sensible heat flux (H)/latent heat flux
(LE)] during the warm season was investigated at 27
sites over 66 site-years within FLUXNET. Variability
in flux partitioning characteristics was large, but some
patterns were detectable. Deciduous forest sites and
the agricultural site had the lowest surface resistance
to water vapor transport (R-c) and B (0.25–0.50).
Coniferous forests had a larger R-c and B (0.50–
1.00 and larger), but R-c and climatological resistance (R-i) varied from year to year. Mediterranean
climate sites usually had the highest net radiation,
R-c, R-i, and B. Grasslands had the largest year-toyear variability, depending mostly on differences in
soil water content and R-c.
Yatskov, M, ME Harmon, and ON Krankina. 2003.
A chronosequence of wood decomposition in the
boreal forests of Russia. Canadian Journal of Forest
Research 33: 1211–1226.
43
For forest ecologists and carbon cycle researchers. The
authors determined specific density of decay classes
and decomposition rates of coarse woody debris from
three hardwood and nine conifer species in Russian
boreal forests. Species differed in annual decomposition rates; rates were highest in Betula pendula and
lowest in Pinus koraiensis and P. siberica. Decomposition rates on the sampled sites were not correlated
with temperature and precipitation, consistent with
other studies in the boreal region.
Integrated Protection of Forests
and Watersheds
Acker, SA, SV Gregory, G Lienkaemper, WA McKee,
FJ Swanson, and SD Miller. 2003. Composition,
complexity, and tree mortality in riparian forests
in the central western Cascades of Oregon. Forest
Ecology and Management 173: 293–298.
Adams, PW. 2002. Assessment and monitoring considerations, pp. 123–131 in National Coastal
Ecosystem Restoration Manual. ORESU-H-02002, Sea Grant Communications, Oregon State
University, Corvallis.
For riparian ecologists, ecologists, and silviculturists. The
authors compared tree composition, stand complexity, and patterns of tree mortality over time in mature
and old-growth stands along low- and middle-order
streams in the western Cascade Range of Oregon. Recruitment of large wood resulting from tree mortality
into stream channels was also assessed. Hardwoods,
Thuja plicata, or both were abundant in stands on
mid-order streams. These streams also had high stand
complexity, as did the upland old-growth stand. In
1996, six of the seven stands studied had exceptionally high tree mortality. This was the year in which the
largest flood during the study took place, but mortality was due primarily to flooding in only one stand, on
an unconstrained reach of a mid-order stream. This
stand yielded a much higher estimated recruitment of
wood than the other stands on mid-order streams.
For landowners, watershed council members, and natural
resource professionals. The author explains the assessment and monitoring of watershed features and functions, including important considerations in comparisons, sampling, and analysis of resource conditions
and management effects.
Adams, PW. 2002. Forests and clean drinking water: Images, rhetoric and reality. Western Forester
47(5): 4–5.
For watershed managers and the general public. The
author discusses how simple images and rhetoric can
distort perceptions and limit understanding of the
relationships between forests and the quality and
quantity of drinking water supplies. Simple riparian
buffer zones and other improved activities can greatly
limit the adverse effects of logging and roads. People
should look at the entire watershed. People are concerned about the 1% of the Eagle Creek basin that
is affected by logging, but the municipal intakes that
are only a mile or two from a major freeway are a far
greater influence on water purity.
Adams, JM, G Piovesan, S Strauss, and S Brown.
2002. The case for genetic engineering of native
and landscape trees against introduced pests and
diseases. Conservation Biology 16: 874–879.
For plant geneticists, forest scientists, and researchers.
Introduced pests and diseases have devastated many
native forest and familiar landscape trees of the
northern temperate zone and are expected to continue devastating more trees. Cautious transfer of
resistance genes from the original host species in the
source region of the pest or disease may undo the
damage. Though there are some problems with the
method, with further work this approach may have
advantages over other techniques such as introgression, including requiring fewer tree generations and
introducing fewer unnecessary genes of nonnative
trees.
Adams, PW, and D Godwin. 2002. Watershed hydrology, pp. II-2.1 to II-2.17 in Watershed Stewardship–A Learning Guide. EM 8714 (revised), Oregon State University Extension Service, Corvallis.
For landowners, watershed council members, and natural
resource professionals. The authors introduce the key
features and functions of watersheds, including the
hydrologic cycle, the role of extreme events, and some
important land use influences.
Arha, K, H Salwasser, and G Achterman. 2003. The
Oregon Plan for Salmon and Watersheds: A Per-
44
Integrated Protection of Forests and Watersheds
could be attributed to water use by vegetation. The
time lag between peak transpiration and minimum
streamflow ranged between about 3 hours in the
early summer to more than 8 hours in the late summer. Only a small fraction of vegetation in the basin
(1–3%) impacted streamflow on this time scale.
spective. INR Policy Paper 2003-03, Institute
for Natural Resources, Oregon State University,
Corvallis.
For policymakers, elected representatives, and natural
resource managers. The Oregon Plan is a state-led
strategy for restoring and conserving native salmonids and the watersheds within which they spend all
or parts of their lives. It encompasses all native salmonids and all watersheds in the state. This paper is a
perspective on the Oregon Plan and what it offers for
state leadership in species conservation. The perspective focuses more on governance matters and less on
implementation through voluntary actions.
Bonin, HL, RP Griffiths, and BA Caldwell. 2003. Nutrient and microbiological characteristics of fine
benthic organic matter in sediment settling ponds.
Freshwater Biology 48: 1117–1126.
For forest scientists, ecosystem managers, and researchers.
Fine benthic organic matter (FBOM) was collected
from settling ponds in catchments in the Pacific
Northwest that contained either old-growth forests
or stands harvested 30 years earlier and replanted.
Forest harvest influenced the chemical characteristics
of FBOM. C:N ratios were significantly higher and
mineralizable N, extractable ammonium, and labile
polysaccharides were all significantly lower in FBOM
from old-growth catchment sediment than in FBOM
from catchments containing harvested stands. Stream
FBOM from harvested basins also appeared more
biodegradable than stream FBOM from old-growth
basins. Most variables differed seasonally in both
the logged and unlogged catchments, indicating that
past timber harvest influenced both composition of
and seasonal fluctuations in FBOM. Comparisons of
the patterns show that settling pond sediments are a
valued surrogate for mountain stream sediments.
Bishaw, B, W Rogers, and W Emmingham. 2002.
Riparian Forest Buffers on Agricultural Lands in the
Oregon Coast Range: Beaver Creek Riparian Project as a Case Study. Research Contribution 38,
Forest Research Laboratory, Oregon State University, Corvallis.
For forest and riparian managers and ecologists. The
Beaver Creek Riparian Buffer Project was established
to develop better information about how to establish
riparian buffers on coastal pastureland near Newport,
Oregon. Intensive site preparation, continued vegetation management, and both fencing and tubing of
tree seedlings were necessary to gain survival and protect seedlings from small rodents, beaver, and cattle.
When cattle were fenced out, the banks of the streams
were protected within a year. Treatments that planted
a single row of trees along the bank had significant
shading 4–7 years after planting. When six rows were
planted, more pasture was taken up, but significant
shading occurred 2–6 years after planting.
Bond, BJ, JA Jones, G Moore, N Phillips, D Post,
and JJ McDonnell. 2002. The zone of vegetation
influence on baseflow revealed by diel patterns of
streamflow and vegetation water use in a headwater basin. Hydrological Processes 16: 1671–1677.
Brasier, CM, DEL Cooke, JM Duncan, and EM Hansen. 2003. Multiple new phenoptypic taxa from
trees and riparian ecosystems in Phytophthora
gonapodyides-P. megasperma ITAS Clade 6, which
tend to be high-temperature tolerant and either
inbreeding or sterile. Mycological Research 107:
277–290.
For tree physiologists and hydrologists. The study compared the temporal dynamics of transpiration and
streamflow over half-hour time increments through
a summer in a small headwater basin. Researchers
concluded that diel (24-hour) patterns in streamflow
45
For forest pathologists. The authors grouped Phytophthora isolates that are associated with Phytophthora
major ITS Clade 6 into eleven taxa on the basis of
phenotype. Three were described morphospecies;
four had been identified but not described, and four
taxa were previously unknown. The authors discuss
the lineages of the isolates. Clade 6 taxa are strongly associated with forest and riparian ecosystems
Integrated Protection of Forests and Watersheds
Photogrammetric Engineering and Remote Sensing
69: 1367–1375.
and tolerate high temperatures. Unlike most Phytophthora clades, they are predominantly inbreeding
or sterile.
For stream and hydrologic modelers and those interested in
photogrammetry and remote sensing. The 30-m digital
elevation models (DEMs) often used in analyzing
aquatic systems have several limitations that may interfere with accurate stream modeling, delineation of
hydrologic units, and slope classification. The Coast
Landscape Analysis and Modeling study (CLAMS)
generated streams, hydrologic units, and slope classes
from 10-m drainage-enforced DEMs and from 30m DEMs. Drainage enforcement was more effective
in improving spatial accuracy of streams and hydrologic unit boundaries than was increased resolution.
Level 2 DEMs generally provided more accurate
delineation of streams and HU boundaries than Level
1. Slopes were better classified by the 10-m DEMs.
Because of these findings, 10-m drainage-enforced
DEMs have been produced for the Coast Range Province of Oregon.
Burns, D, N Plummer, JJ McDonnell, E Busenburg, G
Casile, C Kendall, R Hooper, J Freer, N Peters, K
Beven, and P Schlosser. 2003. The geochemical
evolution of riparian groundwater in a forested
piedmont catchment. Journal of Ground Water 41:
913–925.
For hydrologists and watershed researchers. The authors
studied the principal weathering reactions and their
rates in riparian ground water at the Panola Mountain Research Watershed near Atlanta, Georgia. They
measured chlorofluorocarbons and tritium/helium3 in samples from 19 shallow wells in the riparian
aquifer and 1 borehole in granite to determine the
apparent age of the water. Concentrations of SiO2,
Na+, and Ca2+ were higher downvalley. The age of the
water ranged from 0 to 27 years. Changes in ground
water chemistry were due mostly to the weathering
of plagioclase to kaolinite. Others factors were less
significant. The plagioclase weathering rate appears
similar to the rate calculated in a previous study. The
rate in this study is 3–4 times slower than in other
published laboratory studies.
Cloke, HL, J-P Renaud, AJ Claxton, JJ McDonnell,
MG Anderson, JR Blake, and PD Bates. 2003.
The effect of model configuration on modelled
hillslope-riparian interactions. Journal of Hydrology 279: 167–181.
For hydrologists and riparian researchers. The authors
investigated how model setup decisions are linked to
the consequential behavior of hydrological processes,
using a two-dimensional finite-element model. They
determined that model setup decisions can determine
whether given hillslope processes are present or absent, as well as the size and direction of flux where the
hillslope and riparian area meet. Not exploring these
consequences for given applications can give rise to
misleading inferences about processes.
Chapin, DM, RL Beschta, and HW Shen. 2002. Relationships between flood frequencies and riparian
plant communities in the upper Klamath Basin,
Oregon. Journal of the American Water Resources
Association 38: 603–617.
For plant ecologists and riparian biologists. The authors
investigated how plant communities depend on
infrequent flooding. Plant communities were sampled
along established cross-sections and channel and
floodplain elevations were determined, as were water
surface elevations associated with specific discharges.
At seven of the nine sites, an average peak flow frequency of 4.6 years was required to sustain riparian
plant communities. Return periods >25 years were
needed at the other two. Riparian plant communities
appear to depend strongly on flooding.
Ferguson, BA, TA Dreisbach, CG Parks, GM Filip,
and CL Schmitt. 2003. Coarse-scale population structure of pathogenic Armillaria species
in a mixed-conifer forest in the Blue Mountains
of northeast Oregon. Canadian Journal of Forest
Research 33: 612–623.
Clarke, S, and K Burnett. 2003. Comparison of digital
elevation models for aquatic data development.
46
For silviculturists, forest scientists, and researchers.
Sampling of recently dead or live symptomatic coni-
Integrated Protection of Forests and Watersheds
fers over 16,100 ha of a mixed-conifer forest in the
Blue Mountains of northeast Oregon produced 112
isolates of Armillaria from six tree species. One hundred and eight of the isolates were Armillaria ostoyae
(Romagn.) Herink, and four were North American
Biological Species X (NABS X). Armillaria ostoyae
genet sizes were approximately 20, 95, 195, 260,
and 965 ha, and cumulative colonization of the study
area was at least 9.5%. Estimates of A. ostoyae spread
rate in conifer forests gave age estimates for the genet
ranging from 1,900 to 8,650 years. The authors
discuss possible mechanisms influencing these genets,
the genetic structure and stability of the pathogen,
and implications for disease management.
Freer, J, JJ McDonnell, KJ Beven, NE Peters, DA
Burns, RP Hooper, B Aulenbach, and C Kendall.
2002. The role of bedrock topography on subsurface storm flow. Water Resources Research 38(12):
5-1–5-16.
Filip, GM, LM Ganio, PT Oester, RR Mason, and BE
Wickman. 2002. Ten-year effect of fertilization on
tree growth and mortality associated with Armillaria root disease, fir engravers, dwarf mistletoe, and
western spruce budworm in northeastern Oregon.
Western Journal of Applied Forestry 17: 122–128.
For silviculturists, forest managers, and forest scientists.
Four randomly selected 4-ha plots in a mixed-conifer
forest in northeastern Oregon received one of three
fertilizer treatments or no treatment. After 10 years,
fertilizer had no significant effects on mortality,
diameter growth, vigor, or live crown ratio in grand
fir and western larch or dwarf mistletoe in larch. Fir
mortality was associated with a variety of insects and
diseases, and larch mortality, vigor, and diameter were
associated with dwarf mistletoe. The authors discuss
possible reasons why fertilization in this experiment did not have the same effects as fertilization in
smaller studies.
Filip, GM, SA Fitzgerald, L Yang-Erve, G Laflamme,
JA Beruve, and G Bussieres. 2003. Fire and Armillaria: effects on viability and dynamics in Eastern Oregon, pp. 787–84 in Root and Butt Rot of
Forest Trees. Proceedings of the IUFRO Working
Party 7.02.01, Quebec City, Canada, 16–21 September 2001. Information Report LAU-X-126,
Laurentian Forestry Centre, Canadian Forest
Service, Quebec.
since burning, soil depth, and the presence of an antagonistic fungus, Trichoderma harzianum, on viability
of Armillaria ostoyae colonizing stem segments of red
alder. The stem segments were buried 8 cm or 30 cm
deep in plots that were either burned or not burned.
The antagonistic fungus was buried with half of the
Armillaria-infected segments. One day after burning
in the fall, recovery of A. ostoyae was reduced significantly at the shallower depth, but not at the deeper.
The presence of T. harzianum did not affect Armillaria
recovery immediately but did appear to decrease recovery after several months. Season of burning might
also affect Armillaria recovery.
For forest hydrologists and forest scientists. The authors
conducted a detailed study of subsurface flow and water
table response coupled with digital terrain analysis
(DTA) of surface and subsurface features at the hillslope scale in Panola Mountain Research Watershed
(PMRW), Georgia. Timing, peak flow, recession characteristics, and total flow volume of subsurface storm
flow contributions of macropore and matrix flow in
different sections along an artificial trench face varied
greatly. Where the bedrock surface acts as a relatively
impermeable boundary, local bedrock topography may
be highly significant at the hillslope scale.
Han, H-S, TW Steele, and LD Kellogg. 2003.
DamQuick: A new method for rapidly assessing residual stand damage during partial timber
harvests. Western Journal of Applied Forestry 18:
81–87.
For logging and forestland managers. Partial timber harvest prescriptions are becoming more frequent. The
authors developed and tested a method, DamQuick,
to assess residual stand damage after partial harvesting. The method assesses damage near the extraction
corridors and scales the plot measures to yield stand
level estimates. The method was easy to implement
and provided mean estimates statistically similar to
actual stand damage.
For forest pathologists and silviculturists. The authors
examined the effects of prescribed burning and time
47
Integrated Protection of Forests and Watersheds
ic factors affecting water demand, the authors identified past, current, and projected withdrawal of surface
water in Pacific Coast states of the USA. They also
identified projected demands for recreational uses
that may compete with consumptive uses. To illustrate
the challenges facing water resource managers and
policymakers in designing water allocation policies,
they present a case study of water use issues in the
Klamath Basin.
Hansen, EM, PW Resser, W Sutton, LM Winton, and
N Osterbauer. 2003. First report of A1 mating
type of Phytophthora ramorum in North America.
Plant Disease 87: 1267.
For forest pathologists and mycological geneticists. Phytophthora ramorum was isolated from Viburnum and
Pieris spp. from a nursery in Clackamas County and
from Camellia sp. from a nursery in Jackson County,
Oregon. The microsatellite alleles of the Clackamas
County isolates and European tester isolates were
identical, whereas those of the Jackson County isolates were identical to those of Oregon forest tester
isolates. Thus, the infestations in the two counties
apparently originated separately. The Clackamas
County isolates, which were of A1 mating type,
were received from a Canadian nursery. The Jackson County isolates, which were of A2 mating type,
reportedly originated from a California nursery.
Johnson, GR, BL Gartner, D Maguire, and A
Kanaskie. 2003. Influence of Bravo fungicide applications on wood density and moisture content
of Swiss needle cast affected Douglas-fir trees. Forest Ecology and Management 186: 339–348.
For forest pathologists, wood quality control managers, and
wood anatomists. Wood from trees that had not been
sprayed with Bravo (and were therefore more affected
by Swiss needle cast) had narrower growth rings,
sapwood, and tracheid cell wall thickness and lower
sapwood moisture content that did wood from trees
in adjacent plots that had been sprayed for the 5 years
before the test. The relationships between earlywood
density and earlywood width also were altered in
unsprayed trees, and their latewood proportion and
overall wood density were higher.
Harrington, CA, B Bishaw, and DS DeBell. 2003.
Patterns of survival, damage, and growth for
western white pine in a 16-year-old spacing trial
in western Washington. Western Journal of Applied
Forestry 18: 35–43.
For forest managers, forest scientists, and silviculturists.
The authors planted western white pine trees that
were resistant to white pine blister rust in the early
1980s to determine the effects of spacing on growth.
The planting was in the southern Cascades of Washington. Spacing ranged from 2 to 6 m. In the sixteenth year after planting, the survival rate was about
80%, and 71% of the trees were free of the disease.
Annual growth averaged 0.7 m in height and 1.0 cm
in diameter for the 11th to 16th year. There was damage due to antler rubbing, but this was not associated
with tree spacing. Early growth rates were higher in
the planted stands than in natural stands.
Kasahara, T, and SM Wondzell. 2003. Geomorphic
controls on hyporheic exchange flow in mountain
streams. Water Resources Research 39(1): 3-1–314.
Houston, LL, M Watanabe, JD Kline, and RJ Alig.
2003. Past and future water use in Pacific Coast
States. General Technical Report PNW-GTR588, USDA Forest Service, Pacific Northwest
Research Station, Portland OR.
For hydrologists and water resource managers and planners. On the basis of expected trends in socioeconom-
48
For hydrologists. The influence of stream size and
channel constraint was examined at four stream
reaches. Hyporheic exchange flows were simulated,
using MODFLOW and MODPATH, to estimate
relative effects of channel morphology on the extent of and flow in the hyporheic zone and residence
time of stream water in the zone. Stream size, and
sometimes channel constraint, affected the way in
which channel morphology controlled exchange flows.
Hyporheic exchange in the second-order sites was
driven by pool-step sequences. Pool-riffle sequences
and a channel split created exchange flows with short
residence times, whereas a secondary channel created exchange flows with long residence times. The
fifth-order site was bedrock-constrained and so had
Integrated Protection of Forests and Watersheds
relatively little exchange flow. Models of groundwater flow were helpful in examining the morphologic
features that controlled hyporheic exchange flow.
Channel morphologic features visible on the surface
controlled the development of the hyporheic zone in
the streams.
forest stands showed high rainfall intensities attenuated by canopies. Modeling responses of soil water
in an idealized hillslope to rainfall and throughfall
indicated that the intensity attenuation effect may
be sufficient to prevent loss of slope stability during
large rainstorms. These findings suggest new mechanisms of how forests affect landslide initiation in
steep terrain.
Keim, RF, and AE Skaugset. 2002. Physical aquatic
habitat I. Errors associated with measurement
and estimation of residual pool volumes. North
American Journal of Fisheries Management 22:
145–150.
For forest engineers, forest scientists, and forest managers.
Researchers used precise, digital terrain models of a
third-order mountain stream in Oregon to identify
the sensitivity of measurements to error and the consequences of those errors in calculating the volumes
of residual pools. Researchers found that reach-level
measurements of residual pool volume may be more
appropriate than measurements of individual pools as
metrics in monitoring schemes that use low-precision
measurement techniques.
Kelsey, RG, and G Joseph. 2003. Ethanol in ponderosa pine as an indicator of physiological injury
from fire and its relationship to secondary beetles.
Canadian Journal of Forest Research 33: 870–884.
For entomologists, wildfire ecologists, and forest scientists.
The authors measured ethanol and water in phloem
and sapwood of Pinus ponderosa 16 days after a fire.
Severely scorched trees had 15 times more phloem
ethanol and 53 times more sapwood ethanol than
trees with zero scorch. Ethanol concentrations in
phloem and sapwood were related, as were sapwood
ethanol values at breast height and tree base. Ethanol
accumulated in trees killed in the fire but declined
as surviving trees recovered from their injuries.
Scorched trees attracted beetles more than undamaged trees, indicating ethanol was being released to
the atmosphere. This pattern was stronger in September than in May. Sapwood ethanol in trees with heavy
and severe crown scorch was the strongest predictor
of second-year mortality.
Keim, RF, AE Skaugset, and DS Bateman. 2002.
Physical aquatic habitat II. Pools and cover affected by large woody debris in three western Oregon
streams. Northern American Journal of Fisheries
Management 22: 151–164.
For forest managers, forest engineers, and forest scientists. Large woody debris (LWD), mostly from large
conifers and red alder (Alnus rubra), is important in
affecting stream channel morphology and aquatic
habitat. Researchers added LWD (primarily red alder)
to three third-order streams in the Oregon Coast
Range and used digital terrain models to evaluate physical habitat for salmonids over 3 years. The
results from this study indicate that relatively small
LWD from red alder can modify physical aquatic
habitat effectively.
Mankowski, ME, and JJ Morrell. 2003. Incidence of
Apocephalus horridus in colonies of Camponotus
vicinus and the effect of antibiotic/antimycotic
mixtures on fly emergence (Diptera: Phoridae;
Hymenoptera: Formicidae). Sociobiology 42:
477–484.
Keim, RF, and AE Skaugset. 2003. Modelling effects
of forest canopies on slope stability. Hydrological
Processes 17: 1457–1467.
For researchers in hillslope and forest hydrology. Field
measurements of rainfall and throughfall in two
49
For insect pathologists and parasitologists and wood preservationists. Ant-decapitating flies (Apocephalus horridus) parasitize carpenter ants (Camponotus vicinus).
The authors found flies in three of eight ant colonies
sampled in western Oregon, parasitizing from 1 to
15% of the workers. Medium-sized workers were
more likely to be parasitized than larger ants. Adding propiconazole and tetracycline to a glucose diet
increased ant decapitation, whereas exposing workers
to a temperature of 39 °C for 48 hours decreased
Integrated Protection of Forests and Watersheds
decapitation, apparently affecting development of the
parasites.
Douglas-fir physiology: Net canopy carbon assimilation, needle abscission and growth. Ecological
Modelling 164:211–226.
Manter, DK. 2002. Energy dissipation and photoinhibition in Douglas-fir needles with a fungal-mediated reduction in photosynthetic rates. Journal of
Phytopathology 150: 674–679.
For ecologists and forest scientists. The authors compared the dissipation of absorbed light and potential
for photooxidative damage in Douglas-fir seedlings,
either healthy or infected with Phaeocryptopus gaeumannii. The infection reduced net CO2 assimilation
rates significantly but did not impact chloroplast
pigments significantly. Maximum thermal dissipation
from needles was the same in both types of seedlings.
The infected needles absorbed excess light according
to two experiments, one showing decline in photosystem II (PSII) efficiency and the other showing a lower
photochemical utilization. PSII efficiency was largely
attributable to photooxidative damage, as suggested
by an increase in minimum fluorescence. The findings
indicate an explanation for the greater pathogenicity
of P. gaeumannii in sun-exposed foliage.
Manter, DK, and KL Kavanagh. 2003. Stomatal regulation in Douglas-fir following a fungal-mediated
chronic reduction in leaf area. Trees—Structure
and Function 17: 485–491.
For forest pathologists and tree physiologists. The authors investigated stomatal regulation in Douglas-fir
infected with Swiss needle cast (SNC) to determine
the potential impacts of stomatal occlusion and defoliation on tree productivity. Stomatal conductance
was correlated negatively with fungal colonization
and positively with leaf specific hydraulic conductance. When fungus reduced maximum stomatal
conductance below the potential maximum, stomatal
sensitivity was lower than expected. Losses in productivity resulting from blockage of stomata and needle
loss are compounded by additional losses in leaf
specific hydraulic conductance and reduced stomatal
conductance in still-functioning stomata.
Manter, DK, BJ Bond, KL Kavanagh, JK Stone, and
GM Filip. 2003. Modelling the impacts of the
foliar pathogen, Phaeocryptopus gaeumannii, on
For forest pathologists and physiologists. A model of
stomatal conductance and photosynthesis was parameterized and applied to Douglas-fir trees either
infected or not infected with Swiss needle cast (Phaeocryptopus gaeumannii). Carbon assimilation (Anet)
was highest in the summer and at or below 0 in the
winter, but P. gaeumannii reduced Anet in infected
trees. Total carbon budgets were negative in foliage
with pseudothecia in at least 25% of stomata. The
whole canopy maintained a positive carbon budget
because of the current-year needles that were not
affected by the disease. When the older, more heavily
diseased foliage is abscised shortly after becoming a
carbon sink, there is a mitigating effect on the wholecanopy Anet.
Manter, DK, LM Winton, GM Filip, and JK Stone.
2003. Assessment of Swiss needle cast disease:
Temporal and spatial investigations of fungal colonization and symptom severity. Journal of Phytopathology 151: 344–351.
For ecologists, forest scientists, and mycologists. The authors studied the differences in fungal colonization and
symptom development between north- and south-facing tree plots at three plantations in the Oregon Coast
Range twice a year in 1998 and 1999. Fungal colonization was closely associated with symptom severity.
The three measures of fungal colonization (ergosterol
content, pseudothecia density, and quantitative PCR)
were highly correlated; the ergosterol-pseudothecia relationship was the only one that differed between plots.
Trees growing on warmer slopes had greater levels of
colonization at low to moderate levels of infection, and
plots with a larger amount of solar radiation had more
needle abscission.
McDonnell, JJ. 2003. Where does water go when it
rains? Moving beyond the variable source area
concept of rainfall-runoff response. Hydrological
Processes 17: 1869–1875.
For rainfall-runoff modelers, hydrologists, and riparian
researchers. While tremendous advances have been
50
Integrated Protection of Forests and Watersheds
made in rainfall-runoff modeling, the best models
still rely on the old concept of variable source area
(VSA). The author presents some new ways of viewing
the catchment rainfall-runoff processes that reflect
new data. A catchment can be viewed as a series of
cryptic reservoirs that have coupled unsaturated and
saturated zones, explicit dimensions and porosities,
and connect vertically and laterally in time and space
in linear and non-linear ways. This flexible model
structure, called a box model, better incorporates soft
data and virtual experiments. Also, it may alleviate
the search for scale-invariant processes.
and mass export likely are the relative timing of
riparian and hillslope source contributions and the
connections and disconnections of dominant runoff
areas.
McGlynn, B, J McDonnell, M Stewart, and J Seibert.
2003. On the relationships between catchment
scale and streamwater mean residence time. Hydrological Processes 17: 175–181.
For hydrologists, forest scientists, and riparian researchers. The authors used tritium (3H) to determine the
relationship between landscape characteristics and
streamwater mean residence time (MRT) at the Maimai catchments in New Zealand, a relatively simple
hydrological system. Young waters (<3 years old)
could be estimated because of high precision analysis,
near-natural atmospheric 3H levels, and well-characterized rainfall 3H inputs. MRT did not correlate with
catchment size but was correlated with the median
sub-catchment size of the sampled watersheds. Landscape organization appears to be a first-order control
on MRT.
McGlynn, BL, JJ McDonnell, and DD Bramer. 2002.
A review of the evolving perceptual model of
hillslope flowpaths at the Maimai catchments, New
Zealand. Journal of Hydrology 257: 1–26.
Since the late 1970s, many studies of hillslope flows
have been carried out at the Maimai catchments.
The many approaches and multiple experiments have
provided alternative, sometimes conflicting, interpretations of subsurface water flow that have contributed
to development of a detailed perceptual model of
hillslope hydrology and catchment behavior in steep,
humid catchments. The authors review data sets collected in the catchments and relate them to advances
in understanding of subsurface flow.
McHale, MR, JJ McDonnell, MJ Mitchell, and CP
Cirmo. 2002. A field-based study of soil water and
groundwater nitrate release in an Adirondack forested watershed. Water Resources Research 38(4):
2-1–2-16.
McGlynn, BL, and JJ McDonnell. 2003. Role of discrete landscape units in controlling catchment dissolved organic carbon dynamics. Water Resources
Research 39(4): 3-1–3-18.
For hydrologists, forest scientists, and riparian researchers. The authors investigated the relationship of
storm dissolved-organic-carbon (DOC) dynamics,
catchment landscape units, and catchment scale to
clarify the controls on DOC export dynamics at the
Maimai watersheds in New Zealand. The proportion
of hillslope runoff was smaller on the rising than
on the falling limb of the hydrograph. The proportion of riparian runoff was larger on the rising than
on the falling limb of the hydrograph. There was a
disconnection of hillslope and riparian areas early in
the storm and a reconnection once the hillslope soil
moisture deficits were satisfied. Thus, the first-order
catchment controls on stream DOC concentrations
For hydrologists, forest scientists, and riparian researchers.
A combination of isotopic, chemical, and hydrometric data were used to study sources of stream water
NO3- and the mechanisms delivering NO3- to the
stream in a watershed in the Adirondack Mountains.
During six monitored storms, soil water and till
groundwater dominated stream base flow and storm
flow, with near-stream groundwater and event water
generally contributing little to streamflow. Stream
water NO3- was highest when till groundwater contribution peaked. The authors propose that soil water
and till groundwater mix primarily in hillslope hollows and most stream water NO3- comes from till
groundwater during both base flow and storms.
Minshew, H, PW Adams, JH Huddleston, and D
Godwin. 2002. Watershed soils, erosion, and
conservation, pp. II-4.1 to II-4.39 in Watershed
51
Integrated Protection of Forests and Watersheds
Stewardship—A Learning Guide. EM 8714 (revised). Oregon State University Extension Service,
Corvallis.
pheromone 3-methylcyclohex-2-en-1-one (MCH)
is highly effective in preventing the infestation of
high-risk trees by Douglas-fir beetle, Dendroctonus
pseudotsugae Hopkins. Results of this study indicate
that MCH dispensers eluting 6 mg/d (3X the current
standard rate) and spaced 15 m apart (3X existing
standard distance) can effectively prevent Douglas-fir
beetle infestations.
For landowners, watershed council members, and natural
resource professionals. The authors introduce the key
characteristics and functions of soils and erosion processes in watersheds and some basic erosion control
and watershed enhancement practices.
Rosso, PH, and EM Hansen. 2003. Predicting Swiss
needle cast disease distribution and severity in
young Douglas-fir plantations in coastal Oregon.
Phytopathology 93: 790–798.
Nakamura, F, and FJ Swanson. 2003. Dynamics of
wood in rivers in the context of ecological disturbance, pp. 279–297 in The Ecology and Management of Wood in World Rivers, SV Gregory, KL
Boyer, and AM Gurnell, eds. American Fisheries
Society Symposium 37, American Fisheries Society, Bethesda MD.
For forest pathologists and plantation managers. Swiss
needle cast occurrence was surveyed from Astoria
to Coos Bay, Oregon, to establish the relationship
between disease distribution and environment. The
relationship between disease severity and climate,
topography, soil, and forest stand characteristics was
studied with multiple linear regression and regression
tree. Most of the variability in disease severity was explained by fog occurrence, precipitation, temperature,
elevation, and slope aspect. A disease prediction map
was constructed from the resulting regressing model.
If summer temperatures are relatively low, disease severity appears to increase with warmer, wetter conditions. The model is suitable for hypothesis testing and
for helping in disease management and research.
For ecologists and geomorphologists. The dynamics of
wood in rivers can produce many kinds of ecosystem
disturbance. The authors discuss such disturbances in
a variety of stream types.
Progar, RA, and AR Moldenke. 2002. Insect production from temporary and perennially flowing
headwater streams in western Oregon. Journal of
Freshwater Ecology 17: 391–407.
For ecologists and entomologists. The authors researched
the effect of stream flow on emergent aquatic insect
fauna at three sites in the conifer forests of western
Oregon, using emergence traps. Total density and
biomass of aquatic insects were higher in temporary
streams, but taxonomic richness was higher in perennial streams. The results suggest that the absence of
vertebrate predators permits the arthropod populations in temporary streams to thrive. The streams are
a potential source of colonization and have an important place in the terrestrial food web.
Seibert, J, and JJ McDonnell. 2002. On the dialog
between experimentalist and modeler in catchment hydrology: Use of soft data for multi-criteria
model calibration. Water Resources Research 38
(11): 23-11–23-14.
Ross, DW, GE Daterman, and KE Gibson. 2002.
Elution rate and spacing of antiaggregation
pheromone dispensers for protecting live trees
from Dendroctonus pseudotsugae (Coleoptera:
Scolytidae). Journal of Economic Entomology 95:
778–781.
For research forest entomologists, forest health specialists,
and natural resource managers. The antiaggregation
52
For hydrologists, catchment modelers, and forest engineers. The authors suggest a new modeling method in
which soft data are utilized through fuzzy measures of
model simulation and parameter value acceptability.
They developed a three-box lumped conceptual model
for the Maimai research catchment in New Zealand.
The three-box model fit the parameter values of
runoff (Reff = 0.93) very well, but it did not fit other
criteria, such as new water contributions to peak
runoff. Soft data lowered the Reff values but improved
overall performance. Soft data multicriteria calibration reduced the parameter uncertainty by 60% and
Integrated Protection of Forests and Watersheds
increased model performance. The model developed
represents more real catchment behavior, even though
model efficiencies are lower.
Seibert, J, K Bishop, A Rodhe, and JJ McDonnell.
2003. Groundwater dynamics along a hillslope: A
test of the steady state hypothesis. Water Resources
Research 39(1): 2-1–2-9.
For forest scientists and forest engineers. The authors investigated the usual assumption that the relation between groundwater levels and runoff can be described
as a succession of steady-state conditions by studying
groundwater level data from two opposing hillslopes
along a stream reach in a Swedish till catchment.
Groundwater levels closer than 40 m to the stream
followed the dynamics of the runoff. Farther away
than 40 m, the correlation of groundwater levels and
the dynamics of the runoff were low. Groundwater
levels at equal distances from the stream showed
similar patterns, the correlation increasing with more
distance. This indicates that there is an identifiable
groundwater pattern modeled, but the steady-state
model was not valid for the test site.
Sherlock, MD, JJ McDonnell, DS Curry, and AT Zumbuhl. 2002. Physical controls on septic leachate
movement in the unsaturated zone at the hillslope
scale. Hydrological Processes 16: 2559–2575.
Sherlock, MD, and JJ McDonnell. 2003. A new tool
for hillslope hydrologists: Spatially distributed
groundwater level and soilwater content measured
using electromagnetic induction. Hydrological
Processes 17: 1965–1977.
For hydrologists. Terrain electrical conductivity measurements (EC) derived from multiple electromagnetic induction (EMI) frequencies were compared
with a distributed grid of water-table depth and
soil-moisture measurements in a highly instrumented
hillslope. EC measurements from a low frequency
EMI meter could explain over 80% of the variation
in water table depth across the hillslope. EC measurements with a high frequency meter explained over
70% of soil-moisture variance. EMI technology may
allow collection of many soil- and groundwater depth
measurements with reasonable accuracy.
Skaugset, AE, GH Reeves, and RF Keim. 2002. Landslides, surface erosion, and forest operations in the
Oregon Coast Range, pp. 213–241 in Forest and
Stream Management in the Oregon Coast Range,
SD Hobbs, JP Hayes, RL Johnson, GH Reeves, TA
Spies, JC Tappeiner II, and GE Wells, eds. Oregon
State University Press, Corvallis.
For hydrologists. The authors sought to characterize
water flux within the vadose zone, understand the
physical controls on the flux, and predict how this
ultimately will affect subsurface water quality in a
residential leach field in New York State. Unsaturated
hydraulic conductivity curves were derived from matric potential measurements. A strong upward flux of
soil water occurred between rainstorms. Low matric
potentials were rapidly converted to near-saturated
and saturated conditions after rainfall began. This led
to steep vertical gradients through the near-surface
horizons of the hillslope. The lateral hydraulic gradients were typically 10 times smaller than the vertical,
and flux through the vadose zone was mostly vertical
and short-lived. Soil water remained in the vadose
zone for a long time. Water movement through macropores did not occur on this hillslope.
53
For hydrologists and forest operations managers. Forest management activities, especially timber harvest,
accelerate the natural process of erosion in the forms
of landslides from clearcuts and surface erosion and
landslides from forest roads. Concern about salmonid
species has focused attention on aquatic habitat in
streams draining managed forests. The authors describe the physical setting of the Oregon Coast range,
discuss landslides and surface erosion in unmanaged
and managed Coast Range forests, and treat prevention and mitigation of accelerated erosion in a managed Coast Range forest, with special attention to
best management practices (BMPs). They point out
that BMPs in recent decades have reduced erosion
resulting from forest management. New data indicate
that debris flow from harvest units brings large woody
debris into streams, improving the aquatic habitat.
Developing management strategies that mimic natural timing, quantities, and composition of large woody
debris might be a realistic approach to managing
debris flows.
Integrated Protection of Forests and Watersheds
Theis, WG, and EM Goheen. 2002. Major forest
diseases of the Oregon Coast Range and their
management, pp. 191–212 in Forest and Stream
Management in the Oregon Coast Range, SD
Hobbs, JP Hayes, RL Johnson, GH Reeves, TA
Spies, JC Tappeiner II, and GE Wells, eds. Oregon
State University Press, Corvallis.
For forest pathologists and forest managers. Although
many pathogens occur in coastal forests and are
major disturbance agents, only a few appear to exert
critical effects on the forest ecosystem or on productivity. The authors discuss the basis biology of and
management strategies to deal with several important
root diseases, mistletoe, white pine blister rust, Swiss
needle cast, and common stem decays. They propose a
decision ladder for use in including disease management in management scenarios and suggest several
areas in which research is needed.
scaling. Some questions were answered, but many
challenge the researchers to continue searching.
Vanderbilt, KL, K Lajtha, and FJ Swanson. 2003. Biogeochemistry of unpolluted forested watersheds in
the Oregon Cascades: Temporal patterns of precipitation and stream nitrogen fluxes. Biogeochemistry 62: 87–117.
For ecologists, hydrologists, and forest scientists. The
authors analyzed input and output of organic and inorganic nitrogen (N) in both precipitation and stream
water on six watersheds in the central Cascades. Total
bulk N deposition averaged 1.6–2.0 kg N/ha/year.
Streamwater N export was <1 kg N/ha/year. From
greatest to least, dissolved organic N (DON), particulate organic N (PON), NH4-N, and NO3-N were
the predominant forms of N exported from all watersheds. DON may be related to regional precipitation;
total annual stream discharge was a positive predictor
of annual DON output in all watersheds. Annual discharge did not universally predict the other forms of
N exported. Only DON had consistent seasonal concentration patterns in all watersheds, with its greatest
concentration in November–December. Multiple
biotic controls on nitrate and ammonium concentrations in streams may obscure temporal flux patterns
of dissolved inorganic nitrogen. Using several watersheds from a single climatic zone is important in a
study of this sort.
Thomas, SR, WA Dunstan, B Dell, JM Trappe, and N
Malajczuk. 2003. Pisolithus hypogaeus sp nov.: A
hypogeous representative of the genus Pisolithus
from western Australia. Mycotaxon 87: 405–410
For mycologists. Pisolithus hypogaeus is associated with
eucalyptus ectomycorrhizal hosts in coastal southwestern Australia. It is the first record of a species
in this genus with a hypogeous habit. The authors
describe the relationship of this species to others in
the genus.
Uhlenbrook, S, J McDonnell, and C Leibundgut.
2003. Preface: Runoff generation and implications for river basin modeling. Hydrological Processes 17: 197–198.
Wallin, KF, TE Kolb, KR Skov, and MR Wagner.
2003. Effects of crown scorch on ponderosa pine
resistance to bark beetles in northern Arizona.
Environmental Entomology 32: 652–661.
For forest entomologists, forest managers, and tree
physiologists. A naturally regenerated stand of Pinus
ponderosa was thinned and prescribed burned, and
crown scorch of 40 trees was estimated. Undamaged foliage had a higher net photosynthetic rate in
heavily and severely scorched trees than in trees with
less scorching, especially in the dry season. Constitutive resin volume was negatively related to intensity
of crown scorch in September, although it had not
differed among scorch classes in June. Induced resin
production, however, decreased with increasing scorch
intensity in both months. The proportion of success-
For hydrologists and hydrological modelers. The authors
introduce an issue of Hydrological Processes that consisted of the papers presented at an October 2000
workshop titled Runoff Generation and Implications for
River Basin Modelling. The workshop aimed to bridge
the gap between field-based experimental research
and hydrological basin modeling and to help define
the key state variables controlling runoff generation from headwater catchments to larger basins.
They summarize the questions discussed for process
research, modeling studies, and regionalization and
54
Integrated Protection of Forests and Watersheds
ful colonization by bark beetles (Ips spp. and Dendroctonus spp.) was low all season. Intensity of crown
scorch was generally positively related to colonization
attempts in both pheromone and non-pheromone
treatments.
Steiner. 2003. Nitrate removal effectiveness of a
riparian buffer along a small agricultural stream in
Western Oregon. Journal of Environmental Quality
32: 162–170.
For agriculturists and riparian managers. To determine
the ability of vegetation in riparian zones to reduce
nitrate in water, the authors examined two riparian
zones with similar soils and hydrology but different
vegetation. Drainage water from perennial ryegrass
(Lolium perenne L.) fields in the Willamette Valley ran
into the streams. An uncultivated riparian zone made
up of grasses and herbaceous vegetation can reduce
NO3-N of shallow ground water significantly. Estimates of shallow ground water flow based on Darcy’s
law explain that very little of the streamflow can be
accounted for in this way. Therefore, the potential
for removal of NO3-N was not fully realized. Effective water quality management relies largely on sound
agricultural practices.
Weiler, M, and F Naef. 2003. Simulating surface and
subsurface initiation of macropore flow. Journal of
Hydrology 273: 139–154.
For hydrologists and modelers. The authors simulated
the water flux into macropores at four field sites and
examined the role of macropore drainage area (MDA)
on macropore flow initiation. Individual macropore
MDAs were used to calculate the total relative MDA.
Macropore density primarily controlled for the total
MDA, and surface microtopography strongly influenced its probability distribution. Most macropores
did not receive much water. The simulated probability
distribution of subsurface initiation was more symmetrical and less variable than that derived for surface initiation. Because the amount of water supplied
to each macropore varies, the percolation depth and
transport of solutes in macroporous soils are altered.
Models of infiltration in macroporous soils should
consider these factors.
Wing, MG, and A Skaugset. 2002. Relationships
of channel characteristics, land ownership, and
land use patterns to large woody debris in western
Oregon streams. Canadian Journal of Fisheries and
Aquatic Sciences 59: 796–807.
Weiler, M, BL McGlynn, KJ McGuire, and JJ McDonnell.
2003. How does rainfall become runoff? A combined tracer and runoff transfer function approach.
Water Resources Research 39 (11): 4-1–4-13.
For hydrologists. Hydrographs provide hydrological
data that integrate the variety of ways in which terrestrial runoff is generated with upstream routing. In
the last 2 decades, tracers have supplanted graphical
separation as a more objective way to separate streamflow components. This paper describes a new method
for isotope hydrograph separation, the transfer function hydrograph separation model (TRANSEP), and
illustrates its use with data from two rainfalls at a
catchment at Maimai, New Zealand.
Wigington, PJ, SM Griffith, JA Field, JE Baham,
WR Horwath, J Owen, JH Davis, SC Rain, and JJ
55
For ecologists, hydrologists and land managers. The
authors examined the relationship of channel and
aquatic habitat to the abundance of large woody
debris (LWD) in 3,703 stream reaches drawn from
diverse ownerships, land uses, and land cover types. If
all land uses and covers were included, LWD abundance was related to patterns of ownership and land
use, but the factors decreased in importance when
nonforested land uses were not included. Volume,
frequency, and size of LWD decreased as channel size
increased in forested streams. Stand age and forest distribution were related primarily to LWD size,
rather than volume or abundance. In forested areas,
the geomorphology of stream reaches may be the
most important factor related to LWD. Land managers may want to concentrate on increasing LWD in
larger streams.
Evaluation of Forest Uses,
Practices, and Policy
Adams, DM. 2003. Market and resource impacts of a
Canadian lumber tariff. Journal of Forestry 101(2):
48–52.
Adams, PW. 2002. Taking a stand: SAF position
statements. Western Forester 47(1): 4.
For forest economists and forest policy analysts. A recent
import tariff on Canadian softwood lumber has found
support. Estimates of the traditional market impacts
and resource tradeoffs indicate that gross revenue of
US lumber producers and private timber sellers will
rise, increasing costs to US lumber consumers. Timber harvest will decrease in Canada but increase on
private lands in the US. As a result, softwood inventories will decline.
For members of the Society of American Foresters. Society of American Foresters (SAF) position statements
can be a powerful tool. The statements can be developed at the national, state, division, or chapter level.
Usually the members vote on a position statement,
and, at the national level, the executive committee
must approve them by a two-thirds majority. Once
a statement is developed and issued, it can be used
in newspapers, voter pamphlets, or a variety of other
media to communicate the message.
Adams, DM, and GS Latta. 2003. Private Timber Harvest Potential in Eastern Oregon. Research Contribution 42, Forest Research Laboratory, Oregon
State University, Corvallis.
Ahn, S, AJ Plantinga, and RJ Alig. 2002. Determinants of projections of land use in the South
Central United States. Southern Journal of Applied
Forestry 26: 78–84.
For forest economists, policy analysts, and forest managers. The authors developed projections of future
harvest potentials in eastern Oregon. Even-flow and
market-based projections of future harvest potential
on industrial lands over the next 50 years are approximately half of average harvests over the past 40
years. For NIPF lands, the even-flow projection is
20% higher than the historical harvest average; the
market-based projection indicates potential for a substantial but short-lived increase in near-term harvest.
Inventories on industrial lands rise under both projections, while NIPF inventories remain fairly stable.
Continued loss of land from NIPF to other owners
and uses has limited influence on the market-based
NIPF harvest projection until after 2050. A simulated policy of expanded riparian protection zones
reduces harvest on both ownerships roughly in proportion to the area removed from the harvestable land
base. A simulated requirement to retain 30% more
residual volume in partially cut stands reduces harvest
by 5% on combined private ownerships and increases
total inventory by 13% after 50 years.
For economists and land-use planners. The authors
used an econometric land model in examining historical trends and projecting future uses of land for
agriculture, forestry, and urbanization. If population
increases but stumpage prices do not over the next
50 years, urban and related land uses are expected to
increase, primarily because of conversion of private
timberland. If both population and stumpage prices
increase, both urban/other and private timberland are
projected to increase, whereas agricultural lands will
decrease.
Alexander, SJ, D Pilz, NS Weber, E Brown, and VA
Rockwell. 2002. Mushrooms, trees, and money:
Value estimates of commercial mushrooms and
timber in the Pacific Northwest. Environmental
Management 30: 129–141.
56
For land-use decision makers and forest resource researchers. Four case studies provided much-needed information on the joint production of trees and three species of wild edible mushrooms in different forests of
the Pacific Northwest. The values for timber and for
Evaluation of Forest Uses, Practices, and Policies
mushrooms are site- and species-specific and range
from having the same soil expectation value (SEV) to
the timber having a value 200 times higher. Production economic choices for timber and wild mushrooms are influenced by their relative values, which
in turn are affected by changes in forest management,
yields for mushrooms and trees, and costs.
personnel hurdles and insufficient data with which to
inform the models.
Bettinger, P, DL Johnson, and KN Johnson. 2003.
Spatial forest plan development with ecological
and economic goals. Ecological Modelling 169:
215–236.
For modelers and forest planners and managers. Desired
spatial and temporal future conditions are becoming
important factors in current forest planning goals.
Using a heuristic technique, the authors developed a
spatial forest plan in which habitat for northern spotted owls could be maintained within a certain radius
of a nest at the same time that thinning and group
selection harvest were used to facilitate development of mid- to late-successional forest conditions.
Restraining nesting, roosting, and foraging (NRF)
habitat levels to a minimum of 40% decreased net
present value by nearly 24% and increased average
NRF value by 11% over 100 years. Constraining
habitat levels to a minimum of 80% decreased net
present value by nearly 70% and increased average
NRF by 29%. This planning process allows examination of management options from both economic and
ecological viewpoints.
Bergen, KM, SG Conard, RA Houghton, ES Kasischke, VI Kharuk, ON Krankina, KJ Ranson,
HH Shugart, AI Sukhinin, and RF Treyfeld.
2003. NASA and Russian scientists observe landcover and land-use change and carbon in Russian
forests. Journal of Forestry 101(4): 34–41.
For land managers, land-use policy makers, and forest resource researchers. Several project teams of the NASA
Land-Cover Land-Use Change Program have been
working with Russian organizations to quantify and
understand past, present, and future land-cover and
land-use trends in Russian boreal forests. Results of
the research are discussed in four categories: forest
dynamics, fire and fire behavior, carbon budgets, and
new remote sensing analysis methods. The research
has yielded several positive effects, including new
collaborations making it possible to further knowledge about the influences of land-cover and land-use
changes worldwide.
Bliss, JC. 2003. Sustaining family forests in rural
landscapes: Rationale, challenges, and an illustration from Oregon, U.S.A. Small-scale Forest
Economics, Management, and Policy 1(2): 1–8.
Bettinger, P, and J Sessions. 2003. Spatial forest planning—To adopt, or not to adopt? Journal of Forestry 101(2): 24–29.
For land-use policy decision makers and land managers.
Habitat fragmentation, size of harvest units, cumulative effects of management activities, and other forest
management issues are prompting a new planning approach focusing on the spatial juxtaposition of forest
activities. This approach involves mathematical programming to incorporate all goals into a forest plan.
Spatial forest planning is encouraged by factors such
as regulatory and voluntary guidelines on the patterns
of harvest units and wildlife habitat, the desirability
of efficient forestland use in meeting various goals,
and the need to coordinate activities across several
ownerships in landscape-level plans. It is discouraged by such factors as technological, financial, and
For natural resource professionals. This article synthesizes research on the economic, social, and ecological
roles of family forestlands and makes an argument
for sustaining a mixed-ownership landscape.
Bliss, JC, and AJ Martin. 2003. Nonindustrial private
forests, pp. 221–240 in Introduction to Forest
Ecosystem Science and Management, 3rd ed, R
Young and R Giese, eds. John Wiley & Sons, Inc.,
Hoboken, NJ.
For forestry students. This chapter in an introductory college text surveys the field of nonindustrial private forestry.
Davidson, RA, H Liu, IK Sarpong, P Sparks, and DV
Rosowsky. 2003. Electric power distribution sys-
57
Evaluation of Forest Uses, Practices, and Policies
tem performance in Carolina hurricanes. Natural
Hazards Review 4: 36–45.
For power company researchers and engineers. The
authors investigated power distribution systems in
North and South Carolina during five recent hurricanes. Maximum gust, wind speed, rainfall, and land
cover type were all investigated for their relationship
to power disruption. The results show the effect of the
storm on equipment, how many outages occurred, and
the number of customers affected. Also, the geographic distribution, duration, and causes of outages
are explored.
Emmingham, WH. 2002. Development of ecosystem
management in the Pacific Northwest. Plant Biosystems 136: 167–175.
For ecosystem modelers and forest resource researchers.
Ecosystem management influences forestry practices
differently on federal and private lands. In western
Oregon, most federal land is reserved for noncommodity use. Federal forest managers manage stands
to enhance structural complexity and species diversity
and manage landscapes to imitate natural disturbance. Large industrial owners focus on timber production within legal limits. Some have implemented
Habitat Conservation Plans that increase landscape
planning and diversity. Managers of small private
forests are more likely to define sustainable forestry
in terms of balancing ecological requirements defined
by laws or requirement for green certification against
their aesthetic preferences and need to derive income.
At a landscape level, these differences are sometimes
complementary, providing for different wildlife species.
Edwards, KK, and JC Bliss. 2003. It’s a neighborhood
now: Practicing forestry at the urban fringe. Journal of Forestry 101(3): 6–11.
For forestry professionals. A case study of one watershed at the urban fringe is presented with lessons
drawn for forest managers at the urban-wildland
interface. Stakeholders did not prefer residential
development to actively managed forests. The standards for communications and management that
residents held were higher for corporate and public
forest managers than for individual private managers.
Active opposition to forest management was reduced
when management intentions were effectively communicated and neighbors’ concerns were consistently
acknowledged.
Emmingham, WH. 2002. Status of uneven-aged management in the Pacific Northwest, USA. Forestry
75: 433–436.
For forest managers and forest ecosystem analysts. Classical uneven-aged methods for forest management
developed in Europe have not been widely adopted in
the Pacific Northwest. Instead, foresters are implementing ecosystem management, including practices
such as green tree retention, variable retention thinning, heavy thinning with underplanting, and protection of riparian buffers. Managed uneven-aged forests
will not necessarily develop from these practices.
European experience with uneven-aged management
would benefit forestry in the Pacific Northwest if
managers start with tested basic principles and modify
practices based on research and adaptive management.
Elwood, NE, EN Hansen, and P Oester. 2003. Management plans and Oregon’s NIPF owners: A
survey of attitudes and practices. Western Journal of
Applied Forestry 18: 127–132.
For forest resource researchers and forestry policy analysts.
The authors surveyed nonindustrial private forest
(NIPF) owners in Oregon to determine their demographics and their attitudes toward management
planning. Respondents tended to be older and owned
an average of 208 forested acres on the west side of
the Cascades and 675 forested acres on the east side.
Less than a third of owners had management plans,
and respondents never cited timber production as the
most important ownership objective. Perceived obstacles to developing a plan highlight potential educational opportunities.
Emmingham, WL, P Oester, M Bennett, F Kukulka, K
Conrad, and A Michel. 2002. Comparing shortterm financial aspects of four management options
in Oregon: implications for uneven-aged management. Forestry 75: 489–494.
58
Evaluation of Forest Uses, Practices, and Policies
For forest owners and general public. Four theoretical
management scenarios were set up to compare the
projected 10-year value of timber from the perspective of the typical private family forest owner: hold
for 10 years, thin for even-age, partial cut for uneven-age, and clearcut now. The hold option consistently gave the highest net asset value (NAV) for
timber and land. The thin option was within 2% less,
the partial-cut option averaged about 3% less, and
the clearcut option ranged from 8% to 17% less
than holding. Pine stands of eastern Oregon showed
similar trends, but all other options were within ~6%
of the hold option.
Foster, D, F Swanson, J Aber, I Burker, N Brokaw, D
Tilman, and A Knapp. 2003. The importance
of land-use legacies to ecology and conservation.
BioScience 53: 77.
For ecologists, conservationists, and natural resource policy
makers. Even after a given land use has ceased, it may
continue to influence ecosystem structure and function for decades, or even much longer. Recognizing
these land-use legacies increases our understanding of
modern conditions and reduces mistakes in planning
for the future. The authors consider diverse ecological phenomena and examine terrestrial and aquatic
ecosystems in order to demonstrate how widespread
and important historical land-use influences are in
environmental science and management.
Hobbs, SD, JP Hayes, RL Johnson, GH Reeves, TA
Spies, JC Tappeiner II, and GE Wells. 2002. Moving toward sustainability, pp. 242–259 in Forest and Stream Management in the Oregon Coast
Range, SD Hobbs, JP Hayes, RL Johnson, GH
Reeves, TA Spies, JC Tappeiner II, and GE Wells,
eds. Oregon State University Press, Corvallis.
For resource managers, specialists, technicians, and policymakers. Factors affecting sustainable management
of forest and stream resources in the Oregon Coast
Range are discussed, including integrated resource
management, sustainability, and biological diversity;
changing realities in the Oregon Coast Range; and
challenges facing policymakers.
Huppert, DD, RL Johnson, J Leahy, and K Bell. 2003.
Interactions between human communities and estuaries in the Pacific Northwest: Trends and implication for management. Estuaries 26: 994–1009.
For social scientists, land managers, policy makers, demographers, and economists. The authors examined the
demographics and economies of communities crucial
to the management of estuaries, as well as public
perceptions, attitudes, and values relating to estuarian ecosystems. They found that such communities
are growing more slowly and have an older population
than the rest of the states in which they are located.
Tourism, recreation, and retirement are increasingly
important relative to the extractive natural resource
industries that traditionally have provided the basis of
the local economies. The authors conclude that public
attitudes and values in the communities in the future
will favor stronger environmental protection.
Helvoigt, TL, DM Adams, and AL Ayre. 2003. Employment transitions in Oregon’s wood products
sector during the 1990s. Journal of Forestry
101(4): 42–46.
For forest industry professionals, employment analysts,
and economists. Only 51% of workers who lost jobs
in the wood products sector in Oregon during the
1990s were employed in Oregon by 1998. Almost
half of those who stayed switched to the service or
wholesale-retail trade sectors. As a group, these workers took a pay cut, and their median wage was less
than the median wage for all Oregon workers. The
30% who found employment in Oregon’s northwestern region made 29% more than those who stayed in
the southwestern or eastern portions of the state.
Johnson, RL, and G Stankey. 2002. Forest and stream
management in the Oregon Coast Range: Socioeconomic and policy interactions, pp. 7–30
in Forest and Stream Management in the Oregon
Coast Range, SD Hobbs, JP Hayes, RL Johnson,
GH Reeves, TA Spies, JC Tappeiner II, and GE
Wells, eds. Oregon State University Press, Corvallis.
59
For forest managers, forest economists, and social scientists. The sociopolitical landscape of the Pacific
Northwest is changing, and the changes are expected
Evaluation of Forest Uses, Practices, and Policies
to deeply impact forest management on the Oregon
coast in the current century. The authors discuss
three key dimensions of these changes: how forest resources are conceived of, social acceptability of forest
management practices, and the relationships between
political power and organizational authority.
than other factors, such as country or industry sector,
in the importance of forest certification in marketing.
Karna, J, E Hansen, and H Juslin. 2003. Social responsibility in environmental marketing planning.
European Journal of Marketing 37: 848–872.
For forest products marketers and planners. The authors
describe environmental marketing and discuss how
companies in Europe incorporate social responsibility
in their environmental marketing planning. Corporate social responsibility is defined and the valuebased dimensions are explored.
Joslin, L. 2003. Wilderness information and education in the Three Sisters Wilderness. International
Journal of Wilderness 9(1): 28–29.
For wilderness managers and wilderness management
students. To assess how RE Manning’s emerging principles of wilderness information and education reflect
in the work of the Central Oregon Wilderness Education Partnership (COWEP) through the Deschutes
National Forest, two institutions of higher education,
individual citizen volunteers, and student interns
promote wilderness experience and resource protection in the Three Sisters Wilderness of Oregon.
Kline, J, BJ Butler, and RJ Alig. 2002. Tree planting
in the south: What does the future hold? Southern
Journal of Applied Forestry 26: 99–107.
For natural resource analyst and forest sector researchers. The authors developed models of historical tree
planting in the southern United States and used
the models to compare historical with recent planting trends and to project tree planting 50 years in
the future. Industrial tree planting was projected to
rise gradually with anticipated increasing industrial
harvest, whereas nonindustrial private tree planting
was projected to decline gradually in view of rising
planting costs and less public assistance for tree
planting.
Juslin, H, and E Hansen. 2003. Strategic Marketing
in the Global Forest Industries, Updated Edition,
Authors Academic Press, Corvallis, Oregon.
For forest marketing teachers, professionals, and students. This textbook focuses on the marketing of
forest products. To facilitate the move from the
theoretical to the practical, a unique planning and
modeling approach is taken. Specific illustrations
from companies are used to illustrate key concepts,
and there is an emphasis on marketing globally in
the forest industry.
Kline, JD, DL Azuma, and A Moses. 2003. Modeling
the spatially dynamic distribution of humans in
the Oregon (USA) Coast Range. Landscape Ecology 18: 347–361.
Karna, J, E Hansen, and H Juslin. 2002. Environmental activity and forest certification in marketing
of forest products—a case study in Europe. Silva
Fennica 37: 253–267.
For forest products marketers. The authors surveyed
forest industries and their industrial customers in
Finland, Sweden, Germany, and England with respect
to environmental emphasis and the role of timber
certification in marketing. Most companies have
begun to include environmental issues in their marketing decisions and see forest certification as necessary in marketing forest products. The environmental
activity of the companies had more explanatory power
60
For human ecologists, sociologists, land managers, and
policymakers. Delineating discrete land use categories
when analyzing land use change may be not be appropriate when the social, economic, and ecological
processes being analyzed respond to humans living in
the area. The authors characterize how humans are
distributed and the dynamics of that distribution in
the forest landscape of the Coast Range in western
Oregon. They develop an empirical model describing
the spatial distribution and rate of change in historic
building densities and use it to project changes in
building densities and spatial distribution of buildings. The resulting maps are used as inputs into a
Evaluation of Forest Uses, Practices, and Policies
landscape-level analysis of social, economic, and
ecological processes.
For researchers and analysts in forest and resource economics, conservation biology, and forest policy. Tradeoffs
between biodiversity and timber production are
modeled in three watersheds in the Coast Range of
Oregon. Cost-effective management is compared to
current management under the current configuration of landowners and policies. The first increment
in biodiversity above maximum timber value level is
the most costly. However, biodiversity protection can
be increased at relatively little loss of timber production over much of the feasible range for biodiversity.
Current management strategies appeared to be very
inefficient.
Lach, D, P List, B Steel, and B Shindler. 2003. Advocacy and credibility of ecological scientists in
resource decisionmaking: A regional study. BioScience 53: 170–178.
For all interested in the conduct and communication of science. To explore the extent to which scientists should
act as policy advocates, the authors surveyed four
groups: scientists, resource managers, representatives of interest groups, and members of the involved
public. All groups except the scientists thought an
integrative role for scientists was the best. Scientists preferred an interpretive role. Scientists were
concerned that more active involvement in decision
making could hurt their credibility, especially among
their peers. The other groups considered a scientist
credible if he or she delivered practical results that
they could use.
Marshall, DD, GP Johnson, and DW Hann. 2003.
Crown profile equations for stand-grown western
hemlock trees in northwestern Oregon. Canadian
Journal of Forest Research 33: 2059–2066.
For silviculturists, growth-and-yield modelers, and forest
mensurationists. The authors developed crown profile
equations for western hemlock in northwest Oregon.
Their model divided the crown into upper and lower
portions where the crown width was largest. When
LCW was known, the model explained about 86% of
variation in crown width, but when LCE was predicted from a model the model explained only 66%. The
model is adjustable for other populations of western
hemlock.
Laliberte, AS, and WJ Ripple. 2003. Wildlife encounters by Lewis and Clark: A spatial analysis of interactions between Native Americans and wildlife.
Bioscience 53: 994–1003.
For conservation and restoration biologists and environmental historians. The authors used the entries in the
journals of the Lewis and Clark expedition to assess
the influence of humans on wildlife distribution and
abundance before Euro-American settlement. Species diversity was lower and large mammals were less
abundant in areas with denser human population.
Overhunting before Euro-American settlement,
especially as facilitated by introduction of horses, may
explain why some species present in the archeological record are absent in the historic. Humans may
influence wildlife considerably even when population
densities are relatively low. This influence is often
underestimated when trying to restore conditions to
the state before Euro-American settlement.
McCune, B, SD Berryman, JH Cissel, and AI Gitelman. 2003. Use of a smoother to forecast occurrence of epiphytic lichens under alternative forest
management plans. Ecological Applications 13:
1110–1123.
Lichtenstein, ME, and CA Montgomery. 2003. Biodiversity and timber in the Coast Range of Oregon:
Inside the production possibility frontier. Land
Economics 79: 56–73.
61
For forest managers, lichenologists, and modelers. The
authors forecast how frequently epiphytic lichen
species would appear in a forest landscape, using two
habitat model types: logistic regression and an ecological neighborhood model, SpOcc (Species Occurrence Modeler). They modeled two plans for the Blue
River watershed in the Cascade Range of Oregon,
one a literal application of standard prescriptions in
the Northwest Forest Plant, and the other simulating natural disturbance regimes more closely. Both
models successfully predicted occurrence of Lobaria
Evaluation of Forest Uses, Practices, and Policies
oregana on the basis of elevation and forest structural class. Lobaria was infrequent at elevations about
1000 m and in young, even-aged stands, but was very
similar in frequency in most of the other structural
classes. Differences in elevation and clearcutting affected lichen occurrence much more than retention
levels or, for mature and old-growth stands, structural
class. The strongest models gave similar results for
20 lichen species. The authors recommend use of the
SpOcc model, which has several advantages.
McIver, JD, PW Adams, JA Doyal, ES Drews, BR
Hartsough, LD Kellogg, CG Niwa, R Ottmar, R
Peck, M Taratoot, T Torgersen, and A Youngblood. 2003. Environmental effects and economics of mechanized logging for fuel reduction in
northeastern Oregon mixed-conifer stands. Western Journal of Applied Forestry 18: 238–249.
For forest managers, logging planners, fire reduction
specialists, economists, and ecologists. Two log extraction
systems were used in mixed conifer stands in eastern
Oregon to reduce fuel while protecting residual largediameter western larch, Engelmann spruce, Douglasfir and lodgepole pine. The systems used a single-grip
harvester with either a forwarder or a skyline yarding
system. Reduction in fuel, stem density, and basal
area were similar for the two systems and lowered
fire risk. Thirty-two percent of seedlings and trees
examined showed damage after harvest, with more
damage to residual large trees. Light displacement
of soil, which occurred in 5–43% of the harvested
area, resulted in a short-term increase in soil microarthropods. Compaction, which occurred on 1.4%
of the soil area, decreased abundance of litter microarthropods for at least 1 year. Revenue was similar in
the two systems, but skyline yarding costs were much
higher.
Molina, R, D McKenzie, R Lesher, J Ford, J Alegria,
and R Cutler. 2003. Strategic survey framework
for the Northwest Forest Plan survey and manage
program. General Technical Report PNW-GTR573, USDA Forest Service, Pacific Northwest
Research Station, Portland OR.
tion needs for all Northwest Forest Plan survey and
manage species, setting up and conducting strategic
surveys, and analyzing the information so obtained
for use in the Plan annual species review and in adaptive management. The steps outlined provide guidance for developing priority information, evaluating
and selecting approaches to information gathering,
implementing annual work plans, and managing,
reporting, and transferring information to the annual
species review.
Montgomery, CA. 2002. Compatibility of timber and
conservation: Tracing the tradeoff frontier, pp.
225–232 in Congruent Management of Multiple
Resources: Proceedings from the Wood Compatibility
Initiative Workshop, AC Johnson, RW Haynes, RA
Monserud, eds. General Technical Report PNWGTR-563, USDA Forest Service, Pacific Northwest Research Station, Portland OR.
For researchers and analysts in forest and resource
economics, conservation biology, and forest policy. The
economic approach to modeling compatibility of
multiple forest land uses is described. Two case studies are used to illustrate the approach.
Montgomery, CA. 2002. Ranking the benefits of biodiversity: An exploration of relative value. Journal
of Environmental Management 65: 313–326.
For researchers and analysts in forest and resource
economics, conservation biology, and forest policy. A
survey to explore public attitudes about the multidimensional benefits associated with biodiversity was
administered. A sample of the US population were
asked to rank hypothetical species presented in choice
sets of three species. Ecological benefits outranked
aesthetic, symbolic, recreational, and commodityrelated benefits as a reason to protect wildlife species
from extinction.
Morrell, PD, and JJ Morrell. 2003. Cognitive impact
of a grade school field trip. Journal of Elementary
Science Education 15: 27–36.
For elementary educators. Oregon Wood Magic is designed to provide accurate information about forestry,
and especially about wood products, to third- and
For forest planners, managers, and policy analysts. The
authors describe a process for assessing informa-
62
Evaluation of Forest Uses, Practices, and Policies
fourth-graders. The program combines classroom
lesson plans with a 3-hour field trip to the Wood
Science and Engineering department at Oregon State
University. Student participants were tested before,
within a week after, and 3 months after their field
trip to find out how much they learned and retained.
Their teachers were asked to rate the usefulness of the
lesson plans and the educational value and appropriateness of the activities and demonstrations for the
students. Students acquired significant new information and retained it even after 3 months. Almost all
teachers used at least some lesson plans, and teachers
rated the field trip activities very highly.
deciding to participate in the councils especially cited
stewardship ethic, property rights amid uncertainty,
and action orientation. These results relate to social-psychological reasons for cooperation: perceived
consensus, group identity, and legitimacy of authority.
The findings are helpful where cross-boundary coordination and management are goals.
Schillinger, RR, DM Adams, GS Latta, and AK Van
Nalts. 2003. An analysis of future private timber
supply potential in western Oregon. Western Journal of Applied Forestry 18: 166–174.
For forest managers and economists. Using volume-flow
and market-based models, the authors projected potential timber supply in western Oregon. Their results
indicate that forest industry owners in the area could
continue to cut at recent levels, and levels of harvest
by nonindustrial private forest (NIPF) owners could
rise to near historical peaks. Industrial management
would tend to plantations with thinning and early
density control, whereas NIPF management would
continue to rely on natural regeneration and increase
commercial thinning. Extending the minimum age
of clearcut harvest would cause prices to rise sharply
and harvest to decline in the near term, but depress
prices and increase harvest in the long term for both
types of owners. Owners would lose both income and
land value because of lower prices and restrictions on
harvest timing.
Reichenbach, M, and V Simon-Brown. 2002. Linking
strategic thinking and project planning: The Oregon State University Extension Forestry experience. Journal of Extension 40(4): pp. n.a. Available
at (http://www.joe.org/joe/2002august/iw1.shtml),
last accessed 2/25/05.
For forestry educators and extension foresters. Oregon
State University’s Extension foresters have used a
group-project planning process since the late 1980s.
In 1998, the group-project planning process was
linked to a new strategic thinking process. This enhancement allowed projects to be categorized, prioritized, and implemented based on estimates of time
needed to complete them and time committed to each
goal. Foresters gave the strategic planning process a
6.76 on a scale of 1–10, with individual parts scoring higher. The process links projects to strategic
goals; allows regional, national, and international
issues and trends to influence group projects selected
for implementation; improves working relationships;
and allows new ideas to emerge.
Sessions, J, R Buckman, M Newton, and J Hamann.
2003. The Biscuit Fire: Management Options for
Forest Regeneration, Fire and Insect Risk Reduction
and Timber Salvage. College of Forestry, Oregon
State University, Corvallis.
Rickenbach, MG, and AS Reed. 2002. Cross-boundary cooperation in a watershed context: The sentiments of private forest landowners. Environmental
Management 30: 584–594.
For sociologists, psychologists, and watershed managers.
Fifty qualitative, in-depth interviews were conducted
to determine why nonindustrial private forest landowners do or do not participate in watershed councils,
Oregon’s local, voluntary, collaborative forums created to encourage cross-boundary management. Those
63
For forest managers, economists, and ecologists. At the
request of the Douglas County Board of Commissioners, the authors examined post-fire restoration
considerations for the more than 400,000 acres
burned by the Biscuit fire in southern Oregon. They
examined forest regeneration, fire and insect reduction, and timber salvage with respect to the effects of
1-year, 3-year, and 5-year delays in active management. They conclude that delay will increase risk
of fire and insect damage; adversely affect riparian
habitat, habitat suitable for wildlife that depend on
Evaluation of Forest Uses, Practices, and Policies
old growth, and conditions suitable for conifer regeneration; and result in a rapid decline in economic
value of fire-killed timber.
ing what they already have. The author discusses six
typical organization structures, from the formal to
the very informal. Vision expresses the ideal future;
mission is the responsibilities of the organization;
goals are specific straightforward statements of expectations. Each member of a group has distinct roles
and responsibilities, which are divided into content
and process roles. A successful group will discuss its
hurdles and recognize its achievements.
Shelby, B, J Thompson, M Brunson, and R Johnson.
2003. Changes in scenic quality after harvest–A
decade of ratings for six silviculture treatments.
Journal of Forestry 101(2): 30–35.
For forest engineers and silviculturists. The authors
examined the scenic quality of six forested sites for
11 years. Five forests had been cut in different ways;
the sixth was an old-growth forest. Average ratings
over time were examined with regression analysis. At
the beginning of the study, there was a large variation
of scenic quality, but at the end of the study all the
forests were closer in rating. Scenic quality increased
over the 10 years in all except the patch-cut forest.
Simon-Brown, V. 2002. Effective meetings management, pp. 31–43 in National Coastal Ecosystem
Restoration Manual, S Ridlington, ed. ORESUH-02-002, Sea Grant Publications, Oregon State
University, Corvallis.
For forestry educators, meeting facilitators, and general
public. This chapter highlights the crucial pieces necessary to conduct effective meetings, especially in a
public environment, where tensions can sometimes be
high. Meetings must be fair, open, and honest—the
fair-open-honest triangle. Eleven specific ways to
improve meetings focus on good organization, prior
agreement on ground rules, sensitivity to others, and
good record-keeping. In an effective meeting, people
have clear roles with specific responsibilities. One
effective idea that helps to document agreements and
keep people focused is wall notes.
Shindler, B, and E Toman. 2003. Fuel reduction strategies in forest communities: A longitudinal analysis of public support. Journal of Forestry 101(6):
8–15.
For public relations officers, policy makers, and fire managers. The authors surveyed the same people in 1996
and 2000 regarding their attitudes about fire management programs on federal lands in eastern Oregon
and Washington. They found three major changes
over the four years: more people considered smoke a
problem; citizens considered USDA Forest Service
information programs less reliable than other sources;
and the relationship between the Forest Service and
local residents apparently had deteriorated. Those
responding continued to be in favor of prescribed
fire and mechanized thinning to reduce fuel in local
forests.
Simon-Brown, V, FD Conway, and P Corcoran. 2002.
Process pitfalls: It’s not the science, it’s the people!
Quarterly Newsletter of Association of Leadership
Educators 11(4): 6–10.
Simon-Brown, V. 2002. Choosing your group’s structure, mission, and goals, pp. 17–29 in National
Coastal Ecosystem Restoration Manual, S Ridlington, ed. ORESU-H-02-002, Sea Grant Publications, Oregon State University, Corvallis.
For forestry educators, watershed councils, and general
public. A successful group has clear organizational
structure, roles, responsibilities, mission, and goals.
A group should begin by acknowledging and evaluat-
64
For watershed council members, leadership educators, and
general public. Watershed councils are very important
to riparian restoration, but they have an inherently
difficult design. The authors argue that this design is
not a flaw. The challenge is to work through important issues. Several problems that have been encountered are described and some possible solutions presented. Members should make sure decision-making
processes are stated clearly in writing. The members
must be able to evaluate the group in an open environment. Some concepts must be defined ahead of
time, such as consensus. Conflict must be constructive and attack the problem, not the people. Success
occurs when people can work together effectively.
Evaluation of Forest Uses, Practices, and Policies
Bureau of Land Management were associated with
mature forest cover, and private industry most often
had young forest cover. Nonindustrial private forest
owners had many different types of cover.
Simon-Brown, V, B Withrow-Robinson, M Engle, S
Reed, and S Broussard. 2003. Art as catalyst for
conversation. Women in Natural Resources 24(1):
25–30.
For forest policy makers, forestry educators, and forest
managers. The authors developed an annual exhibition titled “Seeing the Forest: Art about Forests &
Forestry” to connect with sectors of the public that
traditionally have not been reached by extension
foresters. The program was fundamentally educational and sought to engage the public in dialogue.
All types of art have been included in the shows, and
the viewer feedback is as varied as the art itself. Over
three-quarters of the viewers surveyed said the show
increased their understanding of the complexity of
forest issues.
Stankey, GH. 2003. Adaptive management at the
regional scale: breakthrough innovation or mission
impossible? A report on an American experience,
pp. 159–177 in Proceedings of the 2002 Fenner
Conference on the Environment: Agriculture for the
Australian Environment, BP Wilson and A Curtis,
eds. Johnstone Centre, Charles Sturt University,
Albury, Australia.
For forest managers. The 1993 Northwest Forest Plan
established 10 adaptive management areas. The longterm strategy for their management recognized that
experience and knowledge gained over time would
challenge the assumptions that underlay the establishment of the areas. The author presents a literature
review, interviews with program administrators and
university cooperators, and reviews of plans, proposals, and policy papers associated with the adaptive
management areas.
Spies, TA and KN Johnson. 2003. The importance of
scale in assessing the compatibility of forest commodities and biodiversity, pp. 211–235 in Compatible Forest Management, RA Monserud, RW
Haynes, and AD Johnson, eds. Kluwer, Dordrecht,
The Netherlands.
For ecologists, economists, and forest planners and managers. Scale affects how compatibility of biodiversity
and commodities are measured and perceived. The
authors identify major components of scale and
discuss how scale affects how ecological condition is
perceived. They also look at how small-scale management actions scale up to regions with respect to
timber production and biodiversity.
Stankey, GH, BT Bormann, C Ryan, B Shindler,
V Sturtevant, RN Clark, and C Philpot. 2003.
Adaptive management and the Northwest Forest
Plan—Rhetoric and reality. Journal of Forestry
101(1): 40–46.
For forest policy analysts. Adaptive management should
turn management policies into a source of learning
that can improve future actions. For the most part,
implementing adaptive management has failed in the
Northwest Forest Plan. Lack of leadership, increasing
workloads and declining resources that restrict learning-based approaches, and institutional and regulatory environments that prevent innovation are some
of the obstacles to success. The authors suggest that
now is the time to revitalize efforts to make adaptive
management the central strategy of the Northwest
Forest Plan.
Stanfield, BJ, JC Bliss, and TA Spies. 2002. Land
ownership and landscape structure: A spatial
analysis of sixty-six Oregon Coast Range watersheds. Landscape Ecology 17: 685–697.
For landscape ecologists. This paper explores relationships between ownership diversity and forest habitat
diversity in 66 watersheds. As land ownership became increasingly diverse, so did forest cover. Larger
land ownership patches meant larger forest patches.
As connectivity of land ownership increased, so did
connectivity of forest cover. In the watersheds studied, between 29% and 40% of forest cover structure
variability could be explained by land ownership
structure. The USDA Forest Service and the USDI
Walstad, JD, MD Reed, PS Doescher, JB Kauffman,
RF Miller, BA Shindler, and JC Tappeiner. 2003.
Distance education—A new course in wildland
fire ecology. Journal of Forestry 101(7): 16–20.
65
Evaluation of Forest Uses, Practices, and Policies
For natural resource and distance-learning educators. The
authors describe their experience in developing an
interdisciplinary course on wildland fire ecology, to
be offered by distance education. They conclude that
specialists in media production and web-page development are critical to the success of such a project,
which may require $100,000 or more to develop.
They recommend testing the product with a live
student audience. In spite of the challenges, such an
approach to college-level education holds a great deal
of promise and can most likely be financially feasible.
The authors explain how GIS developed and look
at the challenges and opportunities that GIS use is
facing today. Some of the issues are the collection of
data and its accuracy, structure of data, personnel,
legal issues, and certifying of users.
Withrow-Robinson, B, S Broussard, V Simon-Brown,
M Engle, and AS Reed. 2002. Seeing the forest:
Art about forests and forestry. Journal of Forestry
100(8): 8–14.
Wing, MG, and P Bettinger. 2003. An updated
primer on a powerful management tool. Journal of
Forestry 101(4): 4–8.
For forest and natural resource managers and organizations. Geographic information systems (GIS) have had
a great effect on forest and natural resource organizations because they increase mapping and analytical
capabilities, are affordable, and are easily accessible.
66
For forestry educators and interested public. Forestry
comprises many complex related issues that tend to be
oversimplified by the media and public. These extension foresters developed a traveling art show to communicate with new audiences and stimulate dialogue
about these issues. The paintings, photography, and
other art media introduced some science and policy
issues, increased viewers’ awareness of forestry, challenged their beliefs, and stimulated consideration of
other points of view.
Wood Processing and Product
Performance
1,2,3-benzotriazin-4(3H)-one, could also enhance
the MnP pulp bleaching effect.
Bermek, H, K Li, and K-EL Eriksson. 2002. Studies
on inactivation and stabilization of manganese
peroxidase from Trametes versicolor. Progress in
Biotechnology 21: 141–149.
For researchers in forest products and pulp and paper.
High concentrations of H2O2 cause inactivation of
manganese peroxidase (MnP). A method for stabilization of MnP used for pulp bleaching in the
presence of high concentrations of H2O2 has been
developed. Formation of the inactive form of MnP
(Compound III) was shown to be the major reason for
the inactivation in which hydroxyl free radicals or superoxide anion radicals are not involved. In addition,
1-hydroxybenzotriazole, 2,2,6,6-tetramethyl-1-piperidinyloxy free radical, violuric acid, 3-hydroxy1,2,3-benzotriazin-4-(3H)-one and chlorpromazine
stabilized MnP. The stabilizing effect by 1-hydroxybenzotriazole is due to the conversion of the inactive
Compound III to the native enzyme.
Bermek, H, K Li, and K-EL Eriksson. 2002. Studies
on mediators of manganese peroxidase for bleaching of wood pulps. Bioresource Technology 85:
249–252.
For researchers in forest products and pulp and paper. In
order to enhance the bleaching effect of manganese
peroxidase (MnP), unsaturated fatty acids, thiolcontaining compounds, and various other organic
compounds were applied in pulp bleaching experiments with MnP. Thiol-containing compounds did
not improve the pulp bleaching effect by MnP. Some
unsaturated fatty acids, linoleic acid, and linolenic
acid provided a better pulp bleaching effect than did
Tween 80. The correlation between the number of
C=C bonds in a fatty acid and its pulp bleaching
effect was also investigated. The MnP pulp bleaching capability depended on the carboxylic acid used.
A combination of Tween 80 and a carboxylic acid
resulted in higher pulp brightness than that obtained
with Tween 80 alone. A laccase mediator, 3-hydroxy-
Boston, K, and G Murphy. 2003. Value recovery from
two mechanized bucking operations in the southeastern United States. Southern Journal of Applied
Forestry 27: 259–263.
For logging managers. The optimal recovered value of
two mechanized bucking operations, computed by
an individual-stem log optimization program, was
compared with actual recovered value. In one operation, actual value recovery was 95% of optimal; most
of the loss was attributable to loss of potential sawlog volume resulting from faulty calibration of the
machine. In the other operation, only 58% of optimal value was recovered, due mainly to poor length
measurement. Logs were cut too long in the first
operation and too short in the second. Monitoring
mechanized bucking operations with low-cost quality
control methods should help reduce value lost from
inconsistent measurements.
Bouffier, LA, BL Gartner, and JC Domec. 2003. Wood
density and hydraulic properties of ponderosa pine
from the Willamette Valley vs. the Cascade Mountains. Wood and Fiber Science 35: 217–233.
67
For silviculturists, ecologists, and forest and wood scientists. The authors compared ponderosa pine (Pinus
ponderosa) from 30 to 83 years old to determine
differences in wood quality. Four sites were in the
Willamette Valley (WV) and four in the eastern Cascade Mountains (CM). At one of each site type, they
studied the differences in behavior of the wood for
water transport. The WV trees had denser total wood,
earlywood, and latewood between rings 27 and 31
and as distance from the pith increased, but did not
differ from CM trees in latewood proportion or mean
growth ring width. CM trees had less dense total wood
and latewood with increasing ring width; earlywood
density did not change. Trees grown from a CM seed
Wood Processing and Products Performance
source but planted in the WV resembled a WV tree
after the first few rings. Unlike the CM trees, the WV
did not decrease in ks from outer to inner sapwood.
At an applied air pressure of 3.0 MPa, the WV trees
lost 19% of their ks and the CM trees lost 32%. At
4.0 MPa, the CM trees lost more relative water content. Specific conductivity indicated that the WV race
is better adapted to drought.
Bowers, S. 2003. Dichotomous keys for scaling and
grading merchantable quality sawlogs. Western
Journal of Applied Forestry 18: 250–258.
For log scalers and graders. The author developed,
documented, and field-tested a simplified version of
the Westside Grading Guidelines, as published in the
Official Rules Handbook by the Northwest Log Rules
Advisory Group. The study included two dichotomous
keys: a four-step dichotomous key to determine merchantable vs. nonmerchantable logs and an individual
seven-step dichotomous log-grading key. Results were
compared with certified scalers. Participants in the
study measured log length correctly 99% of time. They
recorded scaling diameters correctly at an 89% rate
and merchantability and log grade at 98% and 97%,
respectively.
Chai, Z, KJ Fridley, MO Hunt, and DV Rosowsky.
2002. Creep and creep-recovery models for wood
under high stress levels. Wood and Fiber Science
34: 425–433.
ing, was evaluated with Douglas-fir heartwood and
ponderosa pine sapwood treated with various levels of
fluoride. Ashing limits the typical laboratory to only
eight samples per day. Extraction underestimated
fluoride levels slightly but required less time, allowing
analysis of many more samples.
Dean, TJ, SD Roberts, DW Gilmore, DA Maguire, JN
Long, KL O’Hara, and RS Seymour. 2002. An
evaluation of the uniform stress hypothesis on
stem geometry in selected North American conifers. Trees—Structure and Function 16: 559–568.
For those interested in wood properties and stem structure.
The authors evaluated the uniform stress hypothesis
of stem formation in nine species of conifer by using
regression analysis to compare stem taper to the stem
taper expected if stems distribute stress uniformly
during development. Their results suggest that bending curvature is regulated somewhat by adjustments in
cross sectional area and that stems taper to maintain
a uniform bending curvature. In addition, stem diameter can be predicted from the bending moment when
the modulus of elasticity is relatively constant.
Eiden, C, RJ Leichti, and TH Miller. 2003. Nonlinear dynamic analysis of heavy timber structures
including passive damping devices, p. 44 in Biographies & Abstracts, Forest Products Society 57th
Annual Meeting. June 22–25, 2003, Bellevue WA.
For wood products manufacturers and researchers. This
computational study summarizes lateral forces and
displacements when friction dampers are used in timber frames exposed to seismic ground motion. Some
of the study variables included connection rigidity
and bracing alternatives. Four different structures
with different connection strategies were examined.
Analyzing the results involved three steps: analyzing
without damping effects; evaluating with only semirigid connections, then with rigid connections, and
then with rigid connections and dampers; and analyzing with semi-rigid connections and damping devices
together in the same model.
For wood scientists and wood engineers. Small clear
southern pine specimens were loaded under thirdpoint bending, bent to a stress level between 69% and
91% of the predicted static strength for 34 hours,
and allowed to recover for 1 hour. The authors studied the resulting deflection vs. time behavior. Modified power law functions provide the best fit to both
primary and secondary experimental data. The viscoelastic behavior of wood under high stress levels can
be modeled using this method.
Chen, H, R Rhatigan, and JJ Morrell. 2003. A rapid
method for fluoride analysis of treated wood. Forest Products Journal 53(5): 43–45.
For wood products scientists and manufacturers. Analyzing fluoride in wood by extraction, rather than ash-
Fell, D, EN Hansen, and J Punches. 2002. Segmenting
single-family home builders on a measure of innovativeness. Forest Products Journal 52(6): 28–34.
68
Wood Processing and Products Performance
For forest products managers and general public interested
in forest products and construction. There was explosive growth in popularity of engineered wood in the
1990s, even though these products had been on
the market long before. Two hundred single-family
homebuilders in Washington, Oregon, and California
were surveyed regarding their use of nine innovative building products. The most innovative builders
tended to be larger firms building high-end homes.
were exposed to one of two decay fungi (Gloeophyllum
trabeum or Trametes versicolor) after treatment with 1
of 17 hydroxylamine derivatives. The authors concluded that the potential of the chemicals tested for
preserving wood is low.
Funck, JW, Y Zhong, DA Butler, CC Brunner, and
JB Forrer. 2003. Image segmentation algorithms
applied to wood defect detection. Computers and
Electronics in Agriculture 41: 157–179.
Fell, DR, EN Hansen, and BW Becker. 2003. Measuring innovativeness for the adoption of industrial
products. Industrial Marketing Management 32:
347–353.
For wood products manufacturers and those involved in
wood quality control. Many image segmentation algorithms are used in detection of defects of wood surfaces. Selecting the most appropriate algorithms can
be difficult. The authors used a variety of algorithms
to detect features of wood surfaces and defined measures for examining algorithm performance. The best
performance overall was obtained with a region-based
similarity algorithm that combines clustering and
region-growing techniques. This algorithm was particularly effective with subtle defects. Edge detection
algorithms gave slightly better clearwood detection
and were as effective as the region-growing algorithm
if subtle defects were unimportant. Algorithms and
factors must be chosen in light of the defect detection
requirements of each wood processing application.
For forest industry professionals and market analysts.
There are two traditional methods for measuring
firm innovativeness in industrial markets. One bases
innovativeness on the time in which a single product
is adopted. The other measures the usage of multiple
products at a single point in time. Both the traditional measures fail to capture the time and the degree of
adoption. The authors propose a new innovativeness
measure. The new measure is a hybrid of the other
two measures and captures what the other methods
failed to capture.
Freitag, CM, and JJ Morrell. 2002. Effect of glycol on
movement of borate from fused borate rods. Forest
Products Journal 52(6): 68–74.
For forest scientists, chemists, and forest products managers. Researchers studied the effects of glycol compounds on borate distribution from fused sodium
borate rods, using Douglas-fir heartwood blocks
conditioned to 15%, 30%, or 60% moisture content. The effect of glycol was most pronounced at
30% moisture content. While glycol can slightly
enhance borate diffusion, the added costs probably do
not justify the use of this compound in place of water
as a medium for enhanced diffusion.
Geng, X, and K Li. 2002. Degradation of non-phenolic lignin by the white-rot fungus Pycnoporus
cinnabarinus. Applied Microbiology and Biotechnology 60: 342–346.
For mycologists and wood scientists. The authors studied high molecular weight lignin, both methylated
nonphenolic and unmethylated phenolic, mixed with
softwood pulp and degraded by white rot fungus. After 3 months of incubation, 40% of the methylated
lignin and 70% of the unmethylated were degraded.
Phenolic hydroxyl groups in the lignin greatly increased the degradation rate. Both types of lignin
showed a substantial increase in carboxylic acids after
degradation. The study suggests that a laccase/mediator system is part of the complete degradation of
lignin.
Freitag, C, KC Li, and JJ Morrell. 2003. Potential for
the use of hydroxylamine derivatives as wood preservatives. Forest Products Journal 53(7–8): 77–79.
For wood preservationists and wood products manufacturers. Ovendried blocks of ponderosa pine sapwood
Geng, X, and K Li. 2003. Deinking of recycled mixed
office paper using two endo-glucanases, CelB and
69
Wood Processing and Products Performance
CelE, from the anaerobic fungus Orpinomyces
PC2. TAPPI Journal 2(7): 29–32.
For researchers and pulp and paper manufacturers. The
authors investigated the deinking of recycled mixed
office paper (MOP) with two endo-glucanases, CelB
and CelE. Treatment of MOP pulp with either
enzyme, followed by flotation, significantly reduced
dirt count and residual ink area and increased brightness. CelE appeared to have superior deinking effects
to CelB. CelB was more effective at hydrolyzing pulp
cellulose than CelE. Ink particles in MOP treated
with CelB tended to be smaller than those treated
with CelE; ink particles smaller than 20 µm were not
efficiently removed by flotation.
Geng, X, K Li, IA Kataeva, X Li, and LG Ljungdahl.
2003. Effects of two cellobiohydrolases CbhA and
CelK, from Clostridium thermocellum on deinking
of recycled mixed office paper. Progress in Paper
Recycling 12(3): 6–10.
For researchers in biotechnology and pulp and paper.
Deinking effects of two cellobiohydrolases, CbhA and
CelK, from Clostridium thermocellum were studied on
mixed office paper (MOP). Treatments of MOP with
both CbhA and CelK, followed by flotation, increased
the residual ink area and the number of residual ink
particles. CbhA decreased deinking more, hydrolyzed
cellulose more effectively, and released more reducing
sugars into the pulp slurry than did CelK. A family 3
cellulose-binding domain at the C-terminus of CbhA
likely is important in the higher hydrolytic activity of
CbhA. Treatment of MOP by CbhA and CelK did
not significantly affect the tear and tensile strengths
of handsheets. The inability of both CbhA and CelK
to modify fiber structures substantially is likely to
make both enzymes ineffective for facilitating ink removal from MOP. The interference of both enzymes
on the efficiency of ink removal in the flotation step
may be responsible for reduction of overall deinking
effects by these two enzymes.
Gonzalez-Hernandez, MP, J Karchesy, and EE Starkey. 2003. Research observation: Hydrolyzable
and condensed tannins in plants of northwest
Spain forests. Journal of Range Management 56:
461–465.
For plant biochemists and chemical ecologists. Tannins,
astringent compounds produced by plants, may affect
browsing and grazing by mammals in forest ecosystems. The authors tested for hydrolysable and condensed tannins in 30 plant species grazed by livestock
and deer. Plants in the heather and rose families had
tannins, whereas forbs, grasses and shrubs other than
heathers did not. Rubus sp. had the highest tannin
content. In most species, levels of condensed tannins
were higher than those of hydrolysable tannins. Tannin content changed with the seasons and was highest in the growing season (late winter to early spring,
depending on species).
Gupta, R, and B Wagner. 2002. Effect of metal connector plates on the bending strength of solid sawn
lumber and LVL: A pilot study. Forest Products
Journal 52(10): 71–74.
For wood truss designers. The result of this study may
help dispel the notion that the teeth of metal connector plates damage fiber and weaken its properties.
This study shows that they have no effect on bending
strength. Regular truss lumber was used to determine
the bending strength when a metal connector plate
was affixed to it. In all but five solid-sawn lumber
(SSL) cases, the lumber failed at a knot. The specimens failed between 71 and 332 seconds. In the
laminated veneer lumber (LVL) cases failure usually
occurred at the edge of the plate.
Gupta, R, LR Heck, and TH Miller. 2002. Experimental evaluation of the torsion test for determining
shear strength of structural lumber. Journal of Testing and Evaluation 30: 283–290.
For wood technologists, scientists, and engineers. The
study examines the use of torsion, instead of the standard method of using a shear block, as a test method
to determine shear strength of wood. Shear strength
based on torsion was higher than the shear strength
based on shear block. Shear strength based on torsion
is pure shear strength and, therefore, more representative of shear strength of wood.
Gupta, R, LR Heck, and TH Miller. 2002. Finite-element analysis of the torsion test for determining
shear strength of structural lumber. Journal of Testing and Evaluation 30: 291–302.
70
Wood Processing and Products Performance
For wood technologists, scientists, and engineers. The
study shows that the shear failure in torsion is similar to the shear failure of specimens subjected to
transverse loads. The torsion test is the best practical
method to determine the pure shear strength of fullsize, structural lumber. Uniform shear stress occurs
within the shear span. The shear span begins and ends
at a distance of approximately twice the depth plus
the grip distance away from each end of the specimen.
Shear slippage and the shear-failure plane are parallel
to the grain.
For researchers in forest products and pulp and paper.
A new wood-degrading fungus, Trichophyton rubrum
LKY-7, secretes a high level of laccase in a glucosepeptone liquid medium. Fungal production was barely
induced by 2,5-xylidine. The laccase was purified
to homogeneity with an overall yield of 40%. The
purified laccase had the distinct color and the basic
spectroscopic features of a typical blue laccase, with a
molecular mass of about 65 kDa. Sequencing of the
N-terminus revealed high homology to laccases from
wood-degrading white-rot fungi but little similarity
to laccases from non-wood-degrading fungi. Redox
potential of the enzyme was low, yet it was one of the
most active laccases in oxidizing a series of representative substrates/mediators. Compared with other
fungal laccases, the laccase has a very low Km value
with 2,2’-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid) as a substrate and a very high Km value
with violuric acid as a substrate. Isoelectric point was
4.0. Its optimal pH values were very acidic, although
it was more stable at neutral than at acidic pH. The
laccase oxidized hydroquinone faster than catechol
or pyrogallol; oxidation of tyrosine was not detectable. Sodium azide and sodium fluoride were strong
inhibitors.
Hamilton, ED, DA Butler, and CC Brunner. 2002.
Cutting to order in the rough mill: A sampling approach. Wood and Fiber Science 34: 560–576.
For forest products manufacturers. The authors adapted
an earlier linear programming approach to filling a
cutting order for lumber to handle nonuniform stock
material that contains defects that are not known in
advance of cutting. The new method uses a random
sample to construct a linear program and prices,
rather than cutting patterns, to specify a solution.
The expected cost of filling an order approximately
equals the minimum possible expected cost for sufficiently large order and sample sizes, and there is a
lower bound on the minimum possible expected cost.
Hansen, E, J Seppala, and H Juslin. 2002. Marketing
strategies of softwood sawmills in western North
America. Forest Products Journal 52(10): 19–25.
For sawmill owners and forest product marketers. Data
about the product, customer, market area, and competitive advantage of their marketing strategy were
collected from 52 independent marketing units of
softwood sawmills in Northern California, Oregon,
Washington, and British Columbia. Companies may
not fully understand the forces operating in their
business environment and might need to focus their
strategies better. Further refinement of the theory
underlying competitive strategy and the measurement
of strategy may be needed.
Leavengood, S, and I Petaisto. 2002. WoodDensity.
xls. Wood Products Extension, Corvallis, Oregon.
Available at (http://wood.oregonstate.edu/pubs.
php), last accessed April 15, 2005.
For logging engineers and others interested in estimating
the weights of standing timber. This site provides an
Excel-based program for calculating density of logs
for estimating yields from timber sales. The program
creates the estimate by inputting species, amount of
sapwood, and log size.
Leavengood, S, and Western Wood Products Association. 2003. Lumber Shrinkage Estimator. Western
Wood Products Association, Portland, Oregon.
Available at (http://www.wwpa.org/techguide/
shrinkage.htm), last accessed April 15, 2005.
Jung, H, F Xu, and K Li. 2002. Purification and
characterization of laccase from wood-degrading
fungus Trichophyton rubrum LKY-7. Enzyme and
Microbial Technology 30: 161–168.
For architects, structural engineers, and other users of
western lumber. This site provides an Excel-based program that estimates shrink and swell for western wood
species groups. The user inputs board size, initial
71
Wood Processing and Products Performance
Society 57th Annual Meeting, June 22–25, 2003,
Bellevue WA.
moisture, and final moisture content. The program
can also be used to compare two different species.
For wood products manufacturers and wood products
researchers. Wood quality and potential for production of structural timbers from used utility poles were
investigated. Wood quality appeared to be unaffected
by prior use as a pole. Visual grades from clear wood
properties were compared with actual timbers that are
tested in bending. The wood quality of the samples
was similar to published values. Posts and timbers
with heart center can meet the published allowable
values for No. 1 according to the allowable properties
derived from small clear specimens and the full-size
tests. Visual inspection proved to be a good way to
separate poles that meet the visual grade requirements
for structural products.
Lecourt, M, A Pizzi, and P Humphrey. 2003. Comparison of TMA and ABES as forecasting systems
of wood bonding effectiveness. Holz als Roh- und
Werkstoff 61: 75–76.
For wood products manufacturers and wood products
researchers. The authors tested tannin adhesives
by thermomechanical analysis (TMA) and by an
automated bond evaluation system (ABES). They
conclude that TMA provided greater precision, but
ABES showed faster response to temperature and
hardener variations. There was a clear correlation
between ABES breaking load and TMA modulus of
elasticity (MOE).
Leichti, RJ, K Crews, and RA Hyde. 2002. Meeting
the triple bottom line using a semi-rigid timber
framing system, pp. 362–369 in 7th World Conference on Timber Engineering, Vol. 2, August 12–15,
2002, Shah Alam, Malaysia.
Li, K, and X Xu. 2002. Effects of a cellulose binding domain on deinking of recycled mixed office
paper. Progress in Paper Recycling 11(2): 9–13.
For researchers in biotechnology and pulp and paper. A
fusion protein containing a cellulose binding domain
(CBD) from Cellulomonas fimi endoglucanase A was
prepared and purified. The purified CBD protein was
used in deinking a mixed office paper (MOP). The
deinking process included the repulping of the MOP,
incubation of MOP pulp slurry with the CBD protein, and a flotation.The incubation of the pulp slurry
with the CBD greatly increased total dirt count and
residual ink area. Direct addition of the CBD to the
flotation stage also increased dirt count and residual
ink area significantly. The overall reduction of the
deinking efficiency by the CBD was due to a decrease
in the efficiency of ink removal in the flotation stage.
For forest products researchers and manufacturers, and
builders. The paper argues for care and selection of
materials and systems to meet environmental needs,
as well as special and economic considerations. A
prototype building with a semi-rigid structural system
is described that meets the criteria.
Leichti, RJ, A Tjahyadi, A Bienhaus, R Gupta, T
Miller, and S Duff. 2002. Design and behavior of
friction dampers for two-dimensional braced and
moment-resisting timber frames, pp. 267–274 in
7th World Conference on Timber Engineering, Vol. 3,
August 12–15, 2002, Shah Alam, Malaysia.
For wood products manufacturers and wood products
researchers. Linear and moment-resisting friction
dampers were implemented into connections for
heavy timber frames for seismic energy dissipation.
The design, testing, and energy dissipation characteristics were evaluated.
Li, K. 2003. The role of enzymes and mediators in
lignocellulose degradation, pp. 196–209 in Wood
Deterioration and Preservation, Advances in our
Changing World, B Goodell, D Nicholas, and T
Schultz, eds. ACS Symposium Series 845. American Chemical Society, Washington, DC.
Leichti, RJ, M Meisenzahl, and D Parry. 2003. Structural timbers from retired Douglas-fir utility poles,
p. 45 in Biographies & Abstracts, Forest Products
72
For researchers in forest products and pulp and paper.
White-rot fungi can selectively and efficiently degrade lignin through secreting enzymes. Major lignin-degrading enzymes include lignin peroxidase
Wood Processing and Products Performance
(LiP), manganese peroxidase (MnP) and laccase. LiP
and MnP require hydrogen peroxide (H2O2) for their
activities, but they are inactivated by a high concentration of H2O2. White-rot fungi produce various
H2O2-producing enzymes. LiP oxidizes both phenolic and nonphenolic lignin substructures, whereas
MnP and laccase are only able to degrade phenolic
lignin substructures. An unsaturated fatty acid or a
thio-containing compound enables MnP to degrade
nonphenolic lignin substructures. Several in vitro
studies reveal that some enzymes can synergistically
degrade lignin. As lignin-degrading enzymes are too
bulky to penetrate plant cell walls, the role of various
small organic/inorganic compounds serving as redox
mediators of lignin-degrading enzymes is discussed
in detail.
fir stands: A case study on the Warm Springs
Indian Reservation, Oregon. Western Journal of
Applied Forestry 18: 118–126.
For timber harvesters and harvest methods analysts.
The relationship of logging production and costs
to residual stand damage during commercial thinning was studied by looking at four ground-based
harvesting systems and two harvest unit layout
methods. The costs ranged from $67.75 to $92.66
per thousand board feet. Residual stand damage was
between 20.1% and 62.6%. Equipment size, log
lengths, and layout method affected total residual
stand damage.
Milota, MR. 2002. Drying emissions and environmental regulations, p. 9 in Biographies and Abstracts,
Quality Drying: The Key to Profitable Manufacturing, Sept. 23–25, 2002, Crowne Plaza Montreal
Centre, Montreal, Quebec, Canada.
Liu, Y, and K Li. 2002. Chemical modification of soy
protein for wood adhesives. Macromolecular Rapid
Communications 23: 739–742.
For industry personnel. This report summarizes research at OSU and elsewhere on lumber drying emissions, including factors that affect emissions. Volatile
organic compound emission factors for 22 species
and hazardous air pollutant emission factors for five
species are presented. The new source review, the
boiler panel MACTs, and the regional haze rule, with
the caveat that mill location and proximity to other
company operations can influence how regulations
are applied, are discussed.
For researchers in forest products and wood composites.
Mussel protein is a strong and water-resistant adhesive, but is expensive and not readily available. Soy
protein is inexpensive, abundant, and annually renewable, but suffers from low adhesive strengths and low
water resistances to the bonded products. This study
reveals that introduction of a key functional group
found in the marine adhesive protein to soy protein
could convert the soy protein to a strong and waterresistant wood adhesive.
Mankowski, M, E Hansen, and JJ Morrell. 2002.
Wood pole purchasing, inspection, and maintenance: A survey of utility practices. Forest Products Journal 52(11–12): 43–50.
For utilities, wood treaters, and chemical companies. The
purchasing, inspection, and maintenance practices
related to wood use by 261 North American utilities
were surveyed. The results were compared with a similar survey performed in 1983. The results suggest
the need for more continuing education offerings to
better educate utilities about how to best manage their
wood pole systems for maximum value.
Matzka, PJ, and LD Kellogg. 2003. Harvest system
selection and design for damage reduction in noble
Milota, MR. 2002. Factors affecting lumber kiln VOC
emissions, pp. 4–5 to 4–14 in Proceedings of the
National Council for Air and Stream Improvement,
West Coast Regional Meeting. Portland, OR, September 25–26.
For industry personnel. The author summarizes research at OSU and elsewhere on how temperature,
humidity, and storage affect lumber drying emissions.
Milota, MR. 2003. Designing and implementing an
effective quality programme for softwoods, pp. n.a.
in DryTech 2003: Tools & Technologies to Improve
Kiln Drying operations. Forest Industry Engineering Association, Melbourne, Victoria, Australia,
May 22–23, 2003.
73
Wood Processing and Products Performance
For industry personnel. This paper presents a tutorial
on the factors that contribute to quality kiln drying of
softwoods.
are important to US companies that import radiata
pine.
Milota, MR, and MR Lavery. 2003. Temperature and
humidity effects on emissions of volatile organic
compounds from ponderosa pine lumber. Drying
Technology 21: 165–174.
Milota, MR. 2003. HAP and VOC emissions from
white fir lumber dried at high and conventional
temperatures. Forest Products Journal 53(3):
60–64.
For industry personnel. Emissions were a strong function
of drying temperature for this species. Total hydrocarbon emissions from white fir dried at 115.6 °C
(0.141 kg/m3) were more than double those from the
same wood dried at 82.2 °C (0.052 kg/m3). The high
temperature also caused a 240% increase in methanol
emissions and a 470% increase in formaldehyde emissions. Under Title III of the Clean Air Act Amendment, a large mill could be considered a major source
of pollution when drying at a higher temperature.
Milota, MR. 2003. International developments for
in-kiln systems, pp. n.a. in DryTech 2003: Tools
& Technologies to Improve Kiln Drying Operations.
Forest Industry Engineering Association, Melbourne, Victoria, Australia, May 22–23, 2003
and Rotorua, New Zealand, May 27, 2003.
For industry personnel. Emissions are reported for
ponderosa pine dried throughout the normal range
of kiln conditions. When the temperature increased
from 68 °C to 85 °C, emissions increased 34%.
Total carbon emissions were unaffected by an 11 °C
change in wet-bulb depression. When the lumber
began green and was dried to 12% moisture content,
the average emissions for a drying cycle were 1.4 g/kg
of oven-dry wood.
Moore, JR, DA Maguire, DL Phillips, and CB Halpern. 2002. Effects of varying levels and patterns
of green-tree retention on amount of harvesting
damage. Western Journal of Applied Forestry 17:
202–206.
For tree harvesters, forest managers, and forest resource
analysts. The DEMO (Demonstration of Ecosystem
Management Options) study contains five blocks
with various levels of retention of green trees, either
aggregated or dispersed, and one control block with
100% retention. The authors investigated the fresh
logging wounds on these various blocks. Suppressed
trees tended to sustain the most damage of all the
crown classes. More trees were damaged in the dispersed treatments than the aggregated. The plot that
retained 15% of the trees in a dispersed pattern sustained the most damage. Steeper topography seemed
to be associated with more tree damage.
For industry personnel. This report reviews current systems available for measuring moisture content during
lumber drying.
Milota, MR. 2003. Recent advances in international
kiln drying technologies, pp. n.a. in DryTech
2003: Tools & Technologies to Improve Kiln Drying
Operations. Forest Industry Engineering Association, Melbourne, Victoria, Australia, May 22–23,
2003.
For industry personnel. The author reviews how drying
practices are changing.
Milota, MR. 2003. The U.S. Market—quality demands and consumer concerns, pp. n.a. in DryTech
2003: Tools & Technologies to Improve Kiln Drying
Operations. Forest Industry Engineering Association, Rotorua, New Zealand, May 27, 2003.
Morrell, JJ, D Keefe, and RT Baileys. 2003. Copper,
zinc, and arsenic in soil surrounding Douglas-fir
poles treated with ammoniacal copper zinc arsenate (ACZA). Journal of Environmental Quality 32:
2095–2099.
For industry personnel. This paper is a report to New
Zealand producers about the kiln drying factors that
74
For wood preservationists. The soil surrounding Douglas-fir utility poles treated with ammoniacal copper
zinc arsenate was analyzed to determine levels of copper, zinc, and arsenic. All were elevated in soil next
Wood Processing and Products Performance
to the poles, but declined as soil depth and horizontal
distance from the pole increased. Zinc levels decreased most slowly and arsenic, most quickly. Metals
leaching from ACZA treated poles apparently do not
migrate far in the soil.
Morrell, JJ, A Paillard, D Gnoblei, BL Gartner, MR
Milota, and RG Rhatigan. 2003. Variations in
longitudinal permeability of coastal western hemlock. Wood and Fiber Science 35: 397–400.
For wood preservationists and anatomists, wood products
manufacturers, and wood products researchers. Lumber
of western hemlock varies widely in treatability. The
authors found that air permeability of this species
varied widely among trees and by position within a
tree and tended to decrease with distance from the
bark. The species may have a heartwood zone. The
permeability differences may contribute to the differences in treatability.
companies in New Zealand. Reductions in fleet size
of 25–50% and substantial cost savings appeared
possible, although actual reductions will probably be
smaller because of simplifying assumptions included
in the models. The model is suitable for tactical planning involving small or medium problems, but not
for dispatching log trucks or for use with large operations.
Peshkova, S, and KC Li. 2003. Investigation of chitosan-phenolics systems as wood adhesives. Journal of
Biotechnology 102: 199–207
For wood products researchers, engineers and manufacturers. Pieces of wood veneer only adhere when chitosan,
laccase, and a phenolic compound are all present. In
adhesive systems with a phenolic compound and one
phenolic hydroxyl group, adhesive strength depended
on its chemical structure and its relative oxidation
rate with laccase. Adhesive strengths were correlated
to the viscosity of the adhesive system. If two or three
phenolic hydroxyl groups were adjacent to each other,
there were no correlations. The adhesion mechanisms
of chitosin-phenolic systems may be similar to the
adhesive system of mussels.
Murphy, G. 2003. Procedures for scanning radiata
pine stem dimensions and quality on mechanized
processors. International Journal of Forest Engineering 14(2): 91–101.
For log processors and quality controllers. Productivity,
costs, and value recovery were evaluated for four simulated procedures for scanning and optimal bucking
of pruned and unpruned radiata pine (Pinus radiata): a
conventional scan and three types of automated scan.
The net value recovery was 5–8% higher with the
automated methods than with the conventional scan.
New scanning and optimization equipment on mechanized processors could involve break-even investment costs between $240,000 and $450,000 US.
Peshkova, S, and KC Li. 2003. Investigation of
poly(4-vinylphenol) as a wood adhesive. Wood and
Fiber Science 35: 41–48.
For wood product researchers and developers. The authors
investigated whether a polymer, poly(4-vinylphenol)
(PVP), could be used as a wood adhesive, since it contains components similar to those of environmentally
friendly marine adhesives. An aqueous suspension
of PVP provided up to 3 MPa of shear strength for
wood composites bonded with it. Adding 1,6-hexanediamine or diethylenetriamine to the PVP suspension
increased the strength up to twice that of the aqueous
solution. Curing mechanisms are believed to be the
same for PVP and 1,6-hexanediamine/diethylenetriamine as in the natural quinone-tanning process.
Murphy, G. 2003. Reducing trucks on the road
through optimal route scheduling and shared log
transport services. Southern Journal of Applied
Forestry 27: 198–205.
For forest transportation and site planners. Forest companies are seeking new transport management systems
and novel approaches to using trucking resources
because of concerns about safety, costs, and energy
efficiency. The author developed a model for scheduling truck routes and tested it in two medium-sized
Reeb, JE. 2003. Simulating an extra grader in a sawmill. Forest Products Journal 53(11): 81–84.
75
For wood products manufacturers and managers. Simulations can help managers to anticipate how their deci-
Wood Processing and Products Performance
sions may affect a manufacturing system. They cost
relatively little and cannot damage the system. This
paper describes a simulation experiment done to help
managers of a southern pine sawmill to determine the
effect of adding an extra grader to each of two shifts.
On the basis of the results, an extra grader was added
to each shift.
Reinhold, TA, SD Schiff, DV Rosowsky, and BL Sill.
2002. The case for enhanced in-home protection
from severe winds. ASCE Journal of Architectural
Engineering 8: 60–68.
wildfire. Their distance from markets and resistance
to conventional pressure treatment makes the forests
difficult to use. As the authors suggest, they could
be used as fence posts if an effective treatment was
developed. Through-boring largely did not improve
treatment results, but an ammoniacal-based treatment showed promise.
Rhatigan, RG, MR Milota, JJ Morrell, and MR Lavery. 2003. Effect of high temperature drying on
permeability and treatment of western hemlock
lumber. Forest Products Journal 53(9): 55–58.
For home owners, home builders, and wood engineers.
The authors investigated ways to give protection from
physical wind damage and tornados to houses. The
results were for existing houses but can be extended
to new construction. Tornado hazards discussed are
wind pressures, windborne debris, and falling objects.
The impact resistance of a wall consisting of several
layers can be deduced by adding the impact resistance
of the individual layers. Walls that can resist debris
impact will also tend to resist severe damage due to
tree-fall.
For wood products manufacturers, wood preservationists,
and kiln operators. Western hemlock lumber dried at
high or conventional temperatures showed no difference in gas permeability between the temperature
treatments. Drying at high temperature improved
penetration of ammoniacal copper zinc arsenate, but
reduced penetration of chromated copper arsenate.
Rosowsky, DV. 2002. Performance of timber buildings under high wind loads. Progress in Structural
Engineering and Materials 4: 286–290.
For wood scientists, structural engineers, and home builders. Regions that have the highest risk of hurricanes
and other high wind events often have the most
timber structures. Most low-rise structures, including
residential and light commercial, are made of wood
in countries where wood is used. The author discusses
the implications of high wind for design. Recent and
ongoing research is also highlighted. Also covered
are developments in the areas of codes and standards.
Last, the author surveys the directions of future
research.
Rhatigan, RG, JJ Morrell, and CM Freitag. 2002.
Movement of boron and fluoride from rod formations into Douglas-fir heartwood. Forest Products
Journal 52(11–12): 38–42.
For wood technologists and electric utilities. Using sections of Douglas-fir poles, the authors examined the
ability of boron and fluoride to diffuse from sodium
fluoride and fluoride/boron rods. Chemical levels
in the wood were generally too low to inhibit fungal
attack, regardless of the dosage. The potential for
weakening the poles by drilling additional treatment
holes may outweigh any possible benefits of a higher
dosage.
Rosowsky, DV. 2002. Reliability-based seismic design
of wood shearwalls. ASCE Journal of Structural
Engineering 128: 1439–1453.
Rhatigan, RG, and JJ Morrell. 2003. Use of throughboring to improve CCA or ACZA treatment of
refractory Douglas-fir and grand fir. Forest Products Journal 53(6): 33–35.
For structural engineers and wood engineers. The author
sought to develop a risk-based methodology for seismic design of shear walls based on reliability principles in addition to performance-based concepts. The
study evaluates sources of uncertainty to shear wall
performance, evaluates variables in the peak response,
and develops a risk-based procedure for performance-
For forest managers and wood treatment specialists.
Eastern Oregon forests are full of small-diameter
trees that need to be thinned to prevent catastrophic
76
Wood Processing and Products Performance
based design of wood shear walls. The procedure may
be useful for determining the limits on the seismic
weight to ensure target nonexcedence probabilities for
the different performance levels.
For wood products manufacturers, technologists, and wood
treatment specialists. Pressure within wood caused by
preservative impregnation was examined in Douglasfir (Pseudotsuga menziesii) heartwood and ponderosa
pine (Pinus ponderosa) sapwood. Pressure sensors
attached to sample holders collected the data most
reliably. Pressure reached equilibrium the fastest with
the air treatment in ponderosa pine. In Douglas-fir
heartwood, rapidly changing pressure conditions
probably will affect fluid penetration into the wood
very little.
Rosowsky, D, and S Schiff. 2003. What are our expectations, objectives, and performance requirements
for wood structures in high wind regions? Natural
Hazards Review 4(3): 144–148.
For structural engineers and wood products manufacturers. Tremendous losses of wood-frame structures in
recent storms have spurred considerable efforts to
move from prescriptive design to engineered design
of wood-frame structures built in hurricane-prone
areas, with the assumption that this change will
improve performance of wood-frame constructions in
high winds. Expectations and objectives of the owners, builders, engineers, and insurers are often not in
agreement. The authors draw on their own experience
to discuss some of these incongruencies in expectations, objectives, and performance requirements.
Scott, RJ, RJ Leichti, and TH Miller. 2003. Lateral
force resisting pathways in log structures, p. 44 in
Biographies & Abstracts, Forest Products Society
57th Annual Meeting, June 22–25, 2003, DoubleTree Hotel, Bellevue WA.
For wood products manufacturers and wood products
researchers. This study was conducted to evaluate
lateral force resisting pathways in log structures that
are developed by anchor bolts, thru-rods, friction,
and perpendicular wall connection. To develop a log
wall section that was loaded in shear, a fine-element
model (ANSYS 6.0) was used. The contribution
of construction details to lateral displacement and
capacity was shown. Thru-rod holes can be the source
of increased lateral displacement.
Rosowsky, DV, TG Walsh, and JH Crandell. 2003.
Reliability of residential woodframe construction
from 1900 to present. Forest Products Journal
53(4): 19–28.
For structural engineers, home builders, and wood technologists and engineers. The authors investigated reliability
of selected structural members in single-family dwellings over the last 100 years. Among other things, they
considered the materials, member sizes, framing techniques, and connections schedules of floor joists, roof
rafters, and roof sheathing. Relative risk and a change
in the safety index also were evaluated. Roof rafters
and floor joists have remained relatively constant.
Some decline has occurred due to conservatism used in
assigning nominal design values and reduced strength
resulting from changes in growth characteristics. Reliability of roof sheathing has increased because of better
fastening techniques, dramatically decreasing problems
caused by wind uplift.
Schneider, PF, KL Levien, and JJ Morrell. 2003. Internal pressure measurement techniques and pressure
response in wood during treating processes. Wood
and Fiber Science 35: 282–292.
Singleton, R, DS DeBell, and BL Gartner. 2003.
Effect of extraction on wood density of western
hemlock (Tsuga heterophylla (Raf.) Sarg.) Wood and
Fiber Science 35: 363–369.
For wood products manufacturers, and wood chemists.
Chemical deposits (extractives) in wood samples can
influence wood density, but laboratories that determine wood density may or may not remove them before assessing density. Wood density of young-growth
western hemlock samples was compared before and
after extraction. Ring densities were lower in extracted than in unextracted samples. Extractive levels were
slightly higher near the pith, but extractive content
was generally consistent among samples and along the
radial profile.
Singleton, R, DS DeBell, DD Marshall, and BL Gartner. 2003. Eccentricity and fluting in young-
77
Wood Processing and Products Performance
growth western hemlock in Oregon. Western
Journal of Applied Forestry 18: 221–228.
For wood products manufacturers and tree and stand evaluators. In order to measure eccentricity and fluting,
61 trees from pure western hemlock stands across a
range of age, sites, and densities in the Oregon Coast
Range were cut down. Indices of out-of roundness
(OOR), pith-off-center (POC), and fluting severity
(fluting index) were computed, flutes were counted,
and depth of deepest flute was measured on disks
from breast height, base of live crown, and various
percentage-of-height locations along the bole. Stems
tended to be more out-of-round and have more offcentered piths in the lower portions than in the upper.
Basal and breast-height disks had more, deeper, and
more severe flutes than those from the upper stem.
Tree size (diameter at 4.5 ft) was strongly correlated
with more flute and deeper flutes and higher fluting
severity. OOR was weakly correlated with tree size.
OOR and POC were not significantly correlated with
any other tree or stand attributes. Fluting variables
were correlated positively with stand age and overall tree growth rate and negatively with tree density.
Silvicultural practices involving more rapid growth,
wider spacing, and longer rotations are likely to result
in more extensive fluting, but have little or no effect
on stem eccentricity.
scenario-based and show the vulnerability changes of
the existing building stock because of improvements.
Next, the change in insurance costs can be calculated.
The model shows when retrofitting of existing residential construction is cost-effective; this is called
“zones of economic viability”. Retrofits of up to 40%
of the initial building costs could be economically
viable in some cases.
Tarakandha, B, KS Rao, JJ Morrell, RG Rhatigan,
and DB Thies. 2003. Performance of ACZA- and
CDDC-treated conifers in tropical marine exposures in Krishnapatnam, India. Journal of Tropical
Forest Products 9(1–2): 134–144.
For wood technologists and wood treatment specialists. The
authors assessed the performance of conifer panels
treated with ammoniacal copper zinc arsenate (ACZA)
or copper dimethyldithiocarbamate (CDDC) in marine
environments over 32 months near Krishnapatnam,
India. CDDC did not seem effective for marine exposures under the conditions evaluated, but ACZA appeared to perform well at a retention level of 40 kg/m3.
Taylor, AM, BL Gartner, and JJ Morrell. 2002. Heartwood formation and natural durability: A review.
Wood and Fiber Science 34: 587–611.
For researchers in wood quality and wood products durability. This review builds upon previous reviews of
literature relating to the processes involved in heartwood formation and the natural durability of wood.
It discusses the formation and function of heartwood
in living trees, factors influencing its natural durability, and variations in quantity and quality. Heartwood
forms regularly in tree stems. Heartwood has a different natural decay resistance than sapwood, among
other differences. Understanding heartwood formation is important, including the biochemical processes
through enzymatic analyses.
Smith, B, and E Hansen. 2003. Contributions of
marketing to wood science. Wood and Fiber Science
35: 153.
For wood marketers and scientists. Drawing on the history of forest products marketing, this article argues
that marketing is a science complementary to traditional wood sciences.
Stewart, MG, DV Rosowsky, and Z Huang. 2003.
Hurricane risks and economic viability of
strengthened construction. Natural Hazards Review 4: 12–19.
For economists, insurance companies, and home remodelers. The authors describe a way of determining the
economic viability of costs of changing existing residential structural vulnerability to hurricane damage
compared to cost in insurance losses. The models are
Taylor, AM, BL Gartner, and JJ Morrell. 2003. Coincident variations in growth rate and heartwood
extractive concentration in Douglas-fir. Forest
Ecology and Management 185: 257–260.
For silviculturists and wood anatomists and chemists.
Although extractives can greatly affect heartwood
78
Wood Processing and Products Performance
properties, little is known about heartwood formation and extractive production, nor is there much
information about the effect of environment on
heartwood extractive content. Douglas-fir trees,
53–61 years old, in a recently thinned stand were
sampled and the relationship between growth ring
width and heartwood extractive content was assessed. After thinning, growth ring width increased,
as did extractive content of heartwood thought to
have been formed at the same time. Growth ring
width and extractive content before thinning also
seemed to be roughly correlated. Wood durability
in Douglas fir thus may be affected by silvicultural
treatments affecting growth rate.
Wagner, ER, and EN Hansen. 2002. Methodology for
evaluating green advertising of forest products in
the United States: A content analysis. Forest Products Journal 52(4): 17–23.
For forest products marketers. Environmental (“green”)
advertisements may promote a green life style, present
an image of corporate environmental responsibility,
or indicate a positive relationship between what is
being advertised and the environment. The authors
measured the level of “greenness” over 5 years of
advertisements in magazines and defined five levels,
ranging from extra green to brown. Five to 7% of the
advertisements were classified as extra green or green.
The authors suggested ways to make advertisements
appear greener and noted that companies now focus
advertising mainly on architects.
Thiam, M, MR Milota, and RJ Leichti. 2002. Effect
of high-temperature drying on bending and shear
strength of western hemlock lumber. Forest Products Journal 52(4): 64–68.
For wood scientists and wood products engineers. The authors dried western hemlock (38- by 140-mm) No. 2
and better lumber using conventional and accelerated
kiln schedules and examined the effect of the high
temperature. Conventional drying took 48 hours,
whereas the accelerated schedule required about 24
hours. The drying method did not affect the mean
bending strength, stiffness, or variation associated
with the bending and shear properties. The accelerated method reduced shear strength by 6.4%, base
design bending stress by 8.1%, and base design shear
stress by 14%.
Xu, F, K Li, and TJ Elder. 2002. N-hydroxy mediated
laccase biocatalysis: Recent progress on its mechanism and future prospect of its application. Progress in Biotechnology 21: 89–104.
Thoemen, H, and PE Humphrey. 2003. Modeling the
continuous pressing process for wood-based composites. Wood and Fiber Science 35: 456–468.
For wood products manufacturers. Continuous pressing
has become increasingly important, both quantitatively and qualitatively, over the last 20 years. The
authors developed a simulation model that allows
prediction of temperature, moisture content, air and
water vapor pressure, density profile, adhesive bond
strength, and other important variables. They present
the scientific principles underlying the model and the
boundary conditions and modeling strategy used and
discuss the model predictions for a typical mediumdensity fiberboard production plant.
79
For researchers in biotechnology and pulp and paper. Nhydroxy compounds are important in many biological, pharmacological, and industrial processes. The
traditional emphasis on their metal-ion-chelating
property has recently been shifted to the redox chemistry of the N-OH site, which is of great interest in
developing mediated oxidoreductase-based biocatalyses, such as laccase-catalyzed delignification, decontamination, and organic synthesis. In an N-OHmediated laccase biocatalysis system, N-OH is first
oxidized into N-O• by laccase. A study involving 33
N-OH compounds and seven fungal laccases showed
that the oxidation is controlled by the electron transfer from N-OH to laccase, the rate of which depends
on the redox potential difference between laccase
and N-OH. Higher redox potential tends to reduce
the oxidation rate of N-OH, similar to other laccase
substrates. The redox potential of N-OH is related to
the frontier molecular orbital energy, which is proportional to electron-withdrawing N-phenyl substituents. Balancing the reactivity and stability of N-O•
is key to catalytic efficiency. The prospect of N-OH
mediated laccase biocatalysis is discussed in terms of
applying quantum calculation, rational design, and
methodology development.
Wood Processing and Products Performance
Zhang, C, K Li, and J Simonsen. 2003. A novel woodbinding domain of a wood-plastic coupling agent:
Development and characterization. Journal of Applied Polymer Science 89: 1078–1084.
For researchers in forest products and wood composites.
Poly(N-acryloyl dopamine) (PAD) was successfully
synthesized through free radical homopolymerization
of N-acryloyl-O,O’-diphenylmethyldopamine and
subsequent deprotection. PAD underwent substantial
oxidation and cross-linking reactions at about 80
°C. Therefore, maple veneer samples bonded with
PAD powder at the press temperature of 120 °C
80
had high shear strengths and high water resistance.
In contrast to conventional wood adhesives, PAD
increased, rather than decreased, the shear strengths
of 2-ply laminated maple veneer test specimens after
the specimens had been soaked in water and dried. A
mixture of PAD and polyethylenimine (PEI) resulted
in much higher shear strengths than PAD alone. To
achieve high shear strengths and high water resistances, the maple specimens bonded with PAD/PEI
mixtures had to be cured above 150 °C. Water resistances of the maple specimens bonded with PAD/PEI
mixtures depended on the PDA/PEI weight ratio and
the curing temperature.
Related documents
Download